You are on page 1of 76

ACKNOWLEDGMENTS

Environment Australia wishes to thank the following people for their assistance in
producing this booklet: the principal authors John Johnston, Environment Tasmania,
Gavin Murray, Placer Pacific Limited, with contributions from Dr Ian Ritchie,
Australian Nuclear Science & Technology Organisation; the review team comprising
Graham Terrey, Ian Lambert, Stewart Needham and Dr David Jones; and the steering
committee comprising representatives of the mining industry, government agencies
and peak conservation organisations — the Minerals Council of Australia (MCA), the
Australian Petroleum Exploration Association (APEA), the Australian Institute of
Mining and Metallurgy (AusIMM), individual mining and energy companies,
research institutions, the Australian Conservation Foundation (ACF) and the
Australian Minerals and Energy Environment Foundation (AMEEF). The steering
committee assists the authors without necessarily endorsing their views.

Cover Photo: Kidston Gold Mine Limited, 40 km south of Einasleigh, north


Queensland. Waste dump drainage collection dam is shown with pump back facility
at north dump.

Photo: Placer Pacific Limited

© 1997 Commonwealth of Australia

Information in this document may be reproduced provided that any extracts are fully
acknowledged.

ISBN 0 642 19449 1 of the series 0 642 19418 1

Environment Australia incorporates the environment programs of the Federal


Department of the Environment, Sport and Territories.
FOREWORD
Environment protection is a significant priority for our society. A major role for
government is setting environment standards and ensuring that individuals and
organisations meet them. Increasingly, however, government, industry and
community organisations are working as partners in protecting our environment for
present and future generations.

Representatives of the minerals industry in Australia and Environment Australia (the


environmental arm of the Federal Government) are working together to collect and
present information on a variety of topics that illustrate and explain best practice
environmental management in Australia's mining industry. This publication is one of
a series of booklets aimed at assisting all sectors of the mining industry — minerals,
coal, oil and gas — to protect the environment and reduce the impacts of minerals
production by following the principles of ecologically sustainable development.

These booklets include examples of current best practice in environmental


management in mining from some of the recognised leaders in the Australian
industry. They are practical, cost-effective approaches to environment protection that
exceed the requirements set by regulation.

Australia's better-performing minerals companies have achieved environmental


protection of world standard for effectiveness and efficiency — a standard we want to
encourage throughout the industry in Australia and internationally.

These best practice booklets integrate environmental issues and community concerns
through all phases of mineral production from exploration through construction,
operation and eventual closure. The concept of best practice is simply the best way of
doing things for a given site.

The case studies included in these booklets demonstrate how best practice can be
applied in diverse environments across Australia, while allowing flexibility for
specific sites. Each booklet addresses key issues by presenting:

• basic principles, guidance and advice;


• case studies from leading Australian companies; and
• useful references and checklists.

Mine managers and environmental officers are encouraged to take up the challenge to
continually improve their performance in achieving environment protection and
resource management and apply the principles outlined in these booklets to their
mining operations.

Stewart Needham
Co-Chair, Steering Committee
Assistant Secretary
Office of the Supervising Scientist
Environment Australia

Denis Brooks
Co-Chair, Steering Committee, General Manager-Environment
RGC Limited
EXECUTIVE SUMMARY
The problem of acid drainage, traditionally referred to in Australia as acid mine
drainage (AMD) and in North America as acid rock drainage (ARD), encompasses all
issues associated with the actual and potential environmental effects of sulphide
oxidation resulting from mining activities. Its significant potential for long term
environmental degradation makes it one of the biggest environmental issues facing
the mining industry. Once established in mining wastes such as waste rock stockpiles
or tailings impoundments, acid drainage may persist for tens to hundreds of years and
be difficult and costly to remediate. Increased mobility of dissolved metals is an
associated problem which may occur even when the rate of sulphide oxidation does
not result in net acid generation.

Factors contributing to acid drainage can be classed as primary, secondary and


tertiary. Primary factors are those directly involved in sulphide oxidation - typically,
this occurs when sulphide-bearing rock is exposed to atmospheric oxygen and
moisture, initiating a number of chemical reactions that produce sulphate and acidity.
Secondary factors - such as the presence of other minerals able to neutralise acidity -
consume or alter those products. Tertiary factors are the physical conditions
(materials, minesite topography, climate etc.) that influence the rate of sulphide
oxidation as well as the migration of sulphide oxidation products (that is, sulphate,
metals etc.).

The mechanisms controlling sulphide oxidation and acid generation are not fully
understood. Best practice environmental management for acid drainage is a
developing field in which many questions still remain unanswered. Our understanding
of the problem is being advanced on two fronts through:

• ongoing research into the geochemistry of sulphide oxidation and the factors
influencing oxidation rates; and
• managing and remediating current acid drainage problems in the field.

Best practice environmental management for sulphidic mine wastes requires a risk
management approach. While the primary focus is to prevent acid generation, there is
a hierarchy of appropriate management strategies as follows:

• minimise oxidation rate and isolate higher risk materials from exposure;
• minimise potential for transport of oxidation products from source to receiving
environment; and
• contain and treat acid drainage to minimise risk of significant off-site impacts.

While evaluation of minesite prevention and remediation strategies is an ongoing


process, there is a number of well established best practice principles for minimising
acid drainage. These include early recognition of sulphide oxidation potential and
incorporation of prevention strategies into the various stages of mine planning,
design, operation and closure.

Mine planning to minimise acid drainage is the most cost effective and desirable
solution to the problem. Treatment is less desirable due to the long term nature of acid
drainage and associated high treatment costs.
Environmental management systems for mineral deposits/mining projects containing
sulphide minerals should include, as a high priority, clear objectives and strategies for
dealing with the risk of acid drainage. Mechanisms to monitor and review
performance and, where necessary, revise management strategies, are an essential part
of best practice.

Environmental management of acid drainage requires an understanding of the


geochemistry of the ore body for use in predictive tools. Prediction can be achieved
through static and kinetic testing of mine rock, which can also provide information on
likely metal/sulphate (oxidation products) loads and concentrations. This information
should be incorporated into the mine geological model and mining schedule to
provide for the necessary selective handling and disposal of higher risk materials,
subject to site-specific considerations.

Where acid drainage exists as a result of past mining activities the range of
remediation options may be limited and expensive. Treatment options may be cost
effective under some circumstances. However, due to the long duration of sulphide
oxidation processes, treatment may be required for many years. Prevention and
management at source is the preferred strategy for acid drainage abatement.

The principles of preventing and remediating acid drainage involve a combination of


mechanisms including:

• exclusion of oxygen from sulphidic mine wastes;


• control of water flux within wastes and management of site hydrology to
minimise potential for transport of oxidation products (sulphate, metals etc.);
• neutralisation of acid drainage with alkaline materials; and
• monitoring to characterise wastes and determine the effectiveness of
remediation measures.

The choice of the most appropriate prevention/remediation measures will be affected


by site-specific constraints and the characteristics of the sulphide containing material.
Frequently, innovative and unique strategies are necessary to make the best use of
available resources. This booklet presents a range of strategies and case studies that
demonstrate current approaches to achieving best practice in managing acid drainage
to minimise its associated environmental impact and long term liabilities. Long term
monitoring and evaluation are required to demonstrate the effectiveness of strategies
currently being implemented.

In order to determine the most appropriate solutions for sulphidic mine waste
management at any specific site, a full assessment (cost benefit analysis) of all
potential prevention and remediation options is required prior to implementing the
selected management strategy. In most cases, this may involve lengthy field
evaluation and testing to establish the applicability of each strategy to the particular
site.
CONTENTS

INTRODUCTION

1. WHAT IS ACID DRAINAGE


2. FACTORS INFLUENCING ACID DRAINAGE
2.1 Primary Factors
2.2 Secondary Factors
2.3 Tertiary Factors
3. IMPLICATIONS OF ACID DRAINAGE FOR MINE OPERATORS
4. PREDICTION AND IDENTIFICATION OF ACID DRAINAGE
4.1 Sampling
4.2 Geochemical Static Tests
4.3 Geochemical Kinetic Tests
4.4 Interpretation of Test Results
5. SULPHIDE OXIDATION MANAGEMENT
5.1 Soil Covers
5.2 Water Covers
5.3 Selective Handling and Isolation of Sulphidic Wastes
5.4 Blending
5.5 Bacterial Inhibition
6. TREATMENT STRATEGIES
6.1 Treatment Systems
6.2 Passive Systems
7. MONITORING STRATEGIES

CONCLUSION

REFERENCES AND FURTHER READING

CASE STUDIES
1 Gregory Open Cut Coal Mine, QLD
2 Cadia, waste block modelling, NSW
3 BHP Minerals, Cannington, QLD
4 Placer Pacific Limited
5 Renison Tin Mine, TAS
6 Mt Leyshon, Charters Towers, QLD
7 Clarence Colliery Pty Ltd, Lithgow, NSW
8 Pt Kaltim Prima Coal, Kalimantan, Indonesia
INTRODUCTION
This booklet addresses environmental management to avoid or minimise potential
environmental impacts associated with the oxidation of sulphidic mine wastes
produced by mine excavation and processing. These oxidation processes are
traditionally referred to in Australia as acid mine drainage (AMD) and in North
America as acid rock drainage (ARD). Acid drainage is used in this document to refer
to all issues associated with the actual and potential environmental effects of sulphide
oxidation resulting from mining activities. Acid drainage is one of the most
significant environmental issues facing the mining industry. It affects most sectors of
the industry including coal, precious metals, base metals and uranium.

Experience to date has shown that its effects include acid drainage from waste rock
stockpiles and tailings residue emplacements (impacting on downstream water
quality), development of acid conditions in exposed surface materials (potentially
affecting rehabilitation), increased solubility and/or release of metals (irrespective of
actual pH) and increased salinity or solute loads (oxidation and neutralisation
products).

Ferguson and Erickson (see Further Reading) note that acid drainage "arises from
rapid oxidation of sulphide minerals, and often occurs where such minerals are
exposed to the atmosphere by excavation from the earth's crust. Road cuts, quarries,
or other rock excavations can expose these minerals, but metalliferous mines are a
primary source since economically recoverable metals often occur as orebodies of
concentrated metal sulphides (for example, (iron) pyrite, Fe S2; (copper) chalcopyrite,
CuFeS2; (zinc) sphalerite, ZnS)." The production of acid sulphate soils, a major
problem throughout coastal areas of the world, is also a result of the oxidation of
pyrite in soil of marine and estuarine origins.

1
The environmental consequences of acid drainage can be substantial. For example, in
the United States an estimated 20 000 km of streams and rivers are affected by acid
drainage which has a deleterious effect on aquatic life in streams and in some cases
precludes their beneficial use by other water users. Significant impacts on
groundwater and soil can also occur as a result of mine derived acid drainage.

In 1994 the Canadian Mine Environment Neutral Drainage Program (MEND)


estimated the cost of remediating acid generating mine wastes in Canada to be in
excess of $3 billion.

The impact of acid drainage in Australia has not been fully quantified and the cost of
remediation is unknown. White (1995) estimates the area of acid sulphate soils in
Australia to be 1 000 000 hectares. In early 1996, the Office of the Supervising
Scientist and the Australian Centre for Minesite Rehabilitation Research
commissioned a study of the extent and nature of acid drainage effects from the
minerals industry in Australia. Testing of mine wastes from over sixty coal, gold and
base metal mines by Environmental Geochemistry International (1992) revealed that
up to forty percent of the materials tested were potentially acid forming.

Acid drainage can continue long after mining operations are complete. Historical
mines in the Rio-Tinto region of Spain have been a source of acid drainage for over
2000 years. Abandoned mines can represent a substantial liability for industry and
government once the economic resource at a site has been mined out. High
remediation costs and the persistence of acid generation, once established, results in
acid drainage being a major environmental issue for mine operators, the community
and regulators.

Substantial savings can result where mine planning incorporates strategies for the
management and disposal of acid generating wastes. Addressing the risk of acid
drainage should be an integral part of mine planning and environmental management
systems for the minerals industry (coal, uranium, base metal and precious metal
projects).

Best practice for the management of acid drainage is a developing science and there
are no comprehensive solutions to the problem. Therefore, this booklet addresses best
practice for sulphidic mine waste management from a risk management perspective
and presents a number of case studies highlighting strategies currently being
implemented by the minerals industry. The success of these strategies will depend on
long term evaluation and monitoring as, in many cases, there has been insufficient
time to assess their performance fully.

In addition, research is continuing into the mechanisms influencing oxidation of


sulphide minerals and predictive techniques to provide a more effective risk
management approach. A summary of the extent of current understanding and
management of acid drainage in Australia is provided in the proceedings of the
Second Australian Acid Mine Drainage Workshop held at Charters Towers,
Queensland in March 1995 (see Further Reading).

2
1. WHAT IS ACID DRAINAGE
Oxidation of sulphidic minerals is a natural process resulting from their exposure to
atmospheric conditions. In mining situations this process is accelerated when large
volumes of sulphide rich materials are exposed. Acid drainage resulting from
oxidation of these materials may impact on the immediate and wider environment.

The iron sulphide mineral, pyrite, tends to be the most common sulphide mineral
present in mining situations. Historically, the term acid mine drainage (AMD) has
been applied to the impacts of pyrite oxidation. AMD in coal mining environments is
characterised by low pH (leachate pH values may be as low as 2) and high sulphate
(>2 000 mg/l) and iron. In some circumstances, usually where other metal sulphides
occur, elevated levels of soluble metals may be present in non-acidic effluents.
Therefore, the potential for and nature of acid mine drainage will be site-specific and
a function of the type of mineral deposit.

Typically, acid drainage occurs as runoff or seepages from waste rock stockpiles,
tailings impoundments or coal rejects. Acid drainage may also be discharged from
underground mine workings via adits or shafts or seep from open pit walls where
groundwater is intercepted.

In some cases, as the acid drainage stream is aerated, a characteristically reddish


brown stain forms which discolours drainage channels and stream beds. This results
from the oxidation of ferrous iron in solution which forms a precipitate of insoluble
iron (hydroxide) precipitates.

Photo: John Johnston


Comstock Creek, Mt Lyell, Queenstown, Tasmania.
Shows metal precipitates from acid drainage.

Mining and extraction of sulphide bearing minerals tends to greatly increase the
potential for acid drainage. During mining operations rock material is broken up,
greatly increasing its surface area, and thus exposing sulphide minerals to accelerated
weathering processes. Under these conditions the sulphide minerals oxidise or react
with oxygen and water at varying rates. It is the rate of sulphide oxidation which will
largely determine whether there is a significant potential for acid drainage.

Under the normal pH range of soils and water (pH 5-7), metals released by
weathering of minerals generally precipitate and are relatively immobile. However,
under lower pH conditions, these can remain in solution and be transported off site
3
where they may have a deleterious effect on aquatic ecosystems and other
downstream water users. Impact on groundwater and soils may also occur as a result
of acid drainage discharges.

Acid drainage is not exclusively confined to mine operations and may occur in any
large earthmoving operation which results in the exposure of sulphide minerals, eg
road construction, dam building and airport construction. Oxidation of pyrite is
principally responsible for acid sulphate soils frequently encountered in coastal areas
where soils are developed for agriculture or urban development. Most frequently, acid
sulphate soils develop when low lying coastal areas are drained, lowering the water
table and exposing pyrite in the soil to oxidising conditions.

Acid drainage also occurs naturally in areas of outcropping sulphide bearing rock.
This has previously been used as an exploration tool for mineral deposits where
analysis of stream water samples for sulphate and metals has helped to identify the
location of sulphidic ore bodies.

2. FACTORS INFLUENCING ACID DRAINAGE


A range of physical, chemical and biological processes can influence the generation
of acid drainage and, typically, these factors vary on a site-specific basis. They can be
grouped into primary, secondary and tertiary controls (Ferguson & Erickson). Primary
factors are directly involved in the generation of sulphide oxidation products;
secondary factors consume or alter those products; and tertiary factors are the physical
conditions (materials, minesite topography, climate etc.) that influence the
significance of any sulphide oxidation, the potential for migration into the wider
environment and consumption of oxidation products.

Photo: John Johnston

Mt Lyell, Queenstown, Tasmania. Monitoring of acid drainage leachate from


Magazine Waste Rock Dump. Continuous monitoring records flow, pH and
conductivity. Periodic sampling for metals enables accurate quantification of effluent
load and determination of the success of dump encapsulation in reducing leachate
flow (water ingress) and metal loads (reduced oxygen).

4
2.1 PRIMARY FACTORS

Factors influencing the oxidation of sulphide minerals can be summarised as follows.

• water availability for oxidation and transport;


• oxygen availability;
• physical characteristics of the material;

and to a lesser degree:

• temperature;
• pH;
• ferric/ferrous iron equilibrium; and
• microbiological activity.

Where oxidation occurs in the absence of bacterial catalysts it is known as abiotic and
where bacteria catalyse the reaction it is known as biotic. The oxidation rate of pyrite
is accelerated by the bacteria Thiobacillus ferrooxidans (iron oxidising) and
Thiobacillus thiooxidans (sulphur oxidising), which are associated with nearly all
cases of acid drainage. Maximum oxidation of pyrite occurs between a pH of 2.4 and
3.6, rapidly decreasing above this level. Under acid conditions ferric iron is, in itself,
a powerful oxidising agent which, in turn, may attack other sulphide minerals,
increasing the rate of sulphide oxidation and generation of oxidation products.

Given that water and oxygen are the principal driving factors for AMD it is logical to
isolate the material through various approaches, including encapsulation, as a means
to reduce water and oxygen availability. However research is required to determine
whether the present isolation approaches (such as encapsulation and the thickness and
management of capping materials) provide the desired effect in the long term. There
are measurements which suggest some encapsulation methods may not provide a long
term solution, ie less than 100 years (see Bennett and Pantelis 1991). More
measurements are required to evaluate the long term effectiveness of present isolation,
cover and encapsulation design methods and provide better alternate management
approaches.

Photo: John Johnston


Mt Lyell, Queenstown, Tasmania. Remediation of the Magazine Waste Rock Dump.
Involves compacting clay on the surface of the dump to inhibit the ingress of water
and oxygen. Oxygen permeating the dump enables oxidation of pyritic wastes
producing sulphate, ferrous iron and acidity. Water permeating into the dump
becomes acidified and leaches metals and other elements from the waste rock into
water catchments.
5
2.2 SECONDARY FACTORS

An important factor influencing the acidity generated, and hence pH, is the presence
of other minerals able to neutralise acidity. Carbonates are the only alkaline minerals
which naturally occur in sufficient quantities to be considered effective in the control
and prevention of acid drainage. Silicate minerals and aluminosilicate, such as mica
and clay minerals, have some acid consuming ability but are of minor significance
relative to the carbonates.

The amount of alkaline material in the rock may be sufficient to offset the acid
producing potential of the material and acid drainage will not eventuate as long as the
reaction rates of the respective materials are similar. However, in some cases the
reaction kinetics are such that metals are mobilised even though acid conditions do
not occur. In these situations, sulphate and these metals are indicators of the oxidation
process and may be indicators of impending acid drainage problems.

The pH of the reaction media may also influence the ferrous/ferric iron equilibrium.
At low pH ferric iron acts as an oxidant, while at pH values greater than 3.5 ferric iron
will precipitate as ferric hydroxide Fe(OH)3. Sulphide oxidation rates have also been
shown to increase with increasing partial pressures of oxygen, an effect that is more
pronounced when catalysing bacteria are active.

Photos: Department of Minerals and


Energy (WA)
Teutonic Bore, 270 km north of Kalgoorlie, Western Australia. Acid rock drainage
from lowgrade poly metallic ore/spoil can occur in low rainfall areas (215 mm per
year). Copper sulphate crystals have formed on the surface of ore/spoil.

Photos: Department of Minerals and


Energy (WA)

6
2.3 TERTIARY FACTORS

Climatic regimes can largely determine the impact of acid drainage, the most
significant climatic factors being rainfall and temperature. Moisture is rarely limiting
to the oxidation reaction; however, surface runoff and infiltration resulting from
rainfall are the main mechanisms for transporting oxidation products into receiving
environments. The rate of sulphide oxidation is substantially lower in water than in
air, owing to at least a four orders of magnitude difference in diffusion rates. So
saturation of sulphidic wastes offers a major oxidation control strategy.

In wetter climates, control of acid drainage may permit adequate dilution of reaction
products to be achieved in receiving waters thus minimising the potential effects on
aquatic ecosystems.

In drier climates, while oxidation of sulphide minerals may be occurring in mine


wastes, the hydrological balance is such that there is only minimal transport of
oxidation products. Infrequent rain may result in oxidation products accumulating in
wastes over long dry periods and their release to the environment being controlled by
geochemical processes. However, even in these circumstances the oxidation products
may accumulate at the surface, causing revegetation problems on disturbed land.

Photo: John Johnston


Mt Lyell, Queenstown, Tasmania. Batter slopes covered with compacted clay. Slopes
are a maximum 2.5:1, and 5 m high. Batter design, with runoff drains on each bench,
minimises the potential for erosion of the encapsulating clay layer and allows
placement and retention of a 250 mm revegetation layer of soil material over the
compacted layer on the slope.

The physical nature of mine wastes also affects their acid generating potential. The
rate of acid generation is a function of the surface area of sulphides exposed for
oxidation. Oxidation may be limited to the surface of compacted, fine grained,
weathered stockpiles, compared with stockpiles made up of coarse materials where
wind action (air pressure gradients) and exothermically driven air circulation will
allow oxidation throughout the stockpile. Acid drainage is a function of material

7
characteristics such as particle size, hardness, resistance to weathering and
permeability. These physical characteristics determine oxygen flux and water flux
(percolation rates) through a waste stockpile, which subsequently determines
oxidation and leachate generation rates.

The chemistry of receiving waters may be an important factor in determining the


impact of acid drainage transported into rivers and streams. Metal toxicity in water
can be affected by water hardness or the amount of alkalinity in the water. Where
waters are alkaline this will tend to neutralise acid drainage and promote the
precipitation of metals to less toxic forms. Dissolved organic matter (carbon) may
complex metals in solution, affecting toxicity or bio-availability to aquatic organisms.

3. IMPLICATIONS OF ACID DRAINAGE FOR


MINE OPERATORS
The prevention and control of acid drainage is a major issue for mine operators at
sites where sulphide minerals occur. Ongoing risk assessment of acid drainage
potential and sulphidic waste management during planning, development, operation
and closure of mining developments will result in substantial environmental and
economic benefits.

Inadequate control of acid drainage can result in substantial economic liabilities for
both mine operators and regulators as a result of the long term environmental
degradation. In 1994 the Canadian Mine Environment Neutral Drainage Program
(MEND) estimated the cost of remediating the impacts of acid drainage in Canada to
be in excess of three billion dollars.

Historically the environmental significance of acid drainage was not fully appreciated.
The legacy of many abandoned mine sites producing acid drainage, which is the
source of ongoing environmental damage, highlights the inadequate understanding of
acid drainage in the past. Remediation of abandoned mine sites is often extremely
costly and the cost of acid drainage abatement at older mine sites may exceed the
remaining resource value.

In Australia the best documented example of acid drainage remediation is the Rum
Jungle mine in the Northern Territory. The site was rehabilitated between 1982-86 at
the then cost of $18.6 million*. Rehabilitation objectives were to reduce acid drainage
impacts on the Finniss River, make the site safe to the public and improve the
aesthetic appearance of the mine area. (See Bennet J.W. et al, 1989 in Further
Reading.)

8
Photo: John Johnston
Acid mine drainage from a century old mine adit. Dundas Western Tasmania.

The Mount Lyell Mine in Western Tasmania was the site of continuous copper
mining and processing from 1893 to 1994. A century of mining with little control of
acid drainage has degraded over forty kilometres of rivers and streams which are
consequently unable to support aquatic life of any significance. The Mount Lyell ore
deposit contains in excess of 10 percent sulphide and acidic discharges from the lease
area carries in excess of two tonnes of copper a day into downstream catchments.
Acid drainage is generated as a result of the underground mining operation which
comprises a caving operation beneath a substantial open cut and unconsolidated waste
rock stockpiles on the lease which contain approximately 50 million tonnes of
sulphidic waste (see McQuade et al., OSS Report 104 in Further Reading).

Photo: John Johnston


Mt Lyell mine, Queenstown, Tasmania. Oxygen and temperature monitoring of a
waste rock stockpile by ANSTO.

9
Predictions, based on oxygen and temperature profiles within waste rock stockpiles,
estimate that, in the absence of remediation strategies, oxidation will continue at a
diminishing rate for another six hundred years.
Acid drainage clearly presents a number of potential problems for mine operators:

• impact on mine water quality limiting the reuse of mine water and process water
and creating corrosion problems for mine infrastructure and equipment;
• impact on aquatic ecosystems in downstream environments resulting from
acidity and dissolved metals;
• impact on riparian communities along the downstream drainage channels (for
example, tree deaths);
• possible impact on groundwater quality, particularly shallow aquifers;
• impairment of the beneficial use of waterways downstream of mining operations
for purposes such as stock watering, recreation, fishing, or irrigation;
• difficulties in revegetating and stabilising mine wastes; and
• potential long term liability for mine operators, regulators and the community.

For new developments where initial assessment identifies an acid drainage risk, a
thorough analysis should be undertaken during the feasibility stages of project
evaluation. The acid drainage risk analysis should include:

• characterising the acid generating potential of the materials;


• characterising the mobility of metals and other environmentally significant
constituents of acid drainage;
• estimating the potential for oxidation products to migrate to the environment;
and
• estimating the environmental sensitivity and assimilation capacity of the host
environment.

More recently, mine operators have adopted a range of techniques to identify the
potential for acid drainage and have implemented innovative strategies to control and
minimise the problem. In this way treatment costs and post operational problems
associated with acid drainage can be minimised.
*
Marszalek, A.S. 1996

10
4. PREDICTION AND IDENTIFICATION OF ACID
DRAINAGE
Best practice environmental management can be achieved through the early
recognition of the potential for acid drainage and the adoption of appropriate risk
management strategies. The range of factors affecting sulphide oxidation rates and
subsequent acid drainage must be thoroughly understood so that sound decisions can
be made to minimise impacts of acid generating material disposal practices.

Photo: Renison Limited


CSIRO scientists setting up a bank of piezometers in tailings to study dam hydrology
at Renison Tin Mine, Tasmania.

The first step is to characterise the mine rock types. As well as geological assessment,
this may be facilitated by a number of geochemical tests, broadly classified as static
11
and kinetic tests, which are designed to determine the capacity of particular rock types
to produce acid drainage.

Static testing (Section 4.2) comprises a range of simple, relatively fast and
inexpensive screening tests designed to assess the potential for acid mine drainage in
samples by evaluating the acid generating and acid neutralising processes. Kinetic
tests are used to confirm the findings of static testing, to evaluate the rate of sulphide
oxidation, predict acid drainage characteristics, and to assess potential management
techniques. In the absence of a standardised approach to kinetic testing, interpretation
of results must be done with care and based on professional experience.

Initially, reference should be made to other mining operations in the region,


particularly any mines situated in the same stratigraphic or geological sequence which
may provide some information on the acid generating characteristics of similar ore
bodies and host rocks.

The earliest opportunity to identify the potential for acid drainage at a specific site is
during the exploration stage when drilling is occurring to'prove up' an ore body.

Drill cores can provide information on:


• structure and form of pyrite and associated sulphide minerals such as:
o surface area per unit volume of pyrite in host rocks;
o characteristics of associated minerals and potentially neutralising carbonates;
• percentage of pyrite relative to other potentially neutralising carbonates;
• fabric and texture of rocks, that is, alteration and hardness;
• concentration of metals and elements which may be mobilised by sulphide
oxidation;
• solubility of constituent elements under different environmental conditions; and
• oxidation state of minerals.

Core material can be used for geochemical testing and mineralogical examination
utilising microscopic examination and X-ray diffraction analysis to characterise
sulphide and carbonate species. Sampling for prediction of acid drainage should be
ongoing and may include drill cuttings or rock chip sampling from outcropping rock
units or active mining and development faces.

Hence, the project geologist has a critical role to play in the initial identification and
assessment of acid drainage potential. Relevant information should be logged and
recorded from drill core, and development waste and ore zone core samples retained
for further testing (for example, metallurgy/processing). This information can be
integrated with the geological model for the mineral deposit and form the basis of
initial mine planning strategies.

4.1 SAMPLING

Sampling of drilling products (for example, cuttings, core etc.) to determine sulphide
and carbonate content should be representative, based on accepted statistical
procedures, and similar to the process used to determine other geological
characteristics such as ore grade and reserves.

Representative profiles of all geological units should be sampled. It is important to


identify and quantify low sulphide containing materials and alkaline materials as well

12
as material containing visible sulphides. This information is required so that
consideration can be given to utilisation of acid neutralising material in the
assessment of appropriate management strategies.

The number of samples required and the sampling intensity will be project specific
and depend on a number of factors including:

• geological variability and complexity of rock types; and


• the level of confidence in predictive ability.

Samples should be stored in a cool, dry environment to minimise sulphide oxidation


prior to testing. Geochemical static tests may require as little as 2 grams of sample
and kinetic testing a minimum of 500 grams. It is preferable to collect more sample
than required for testing to allow for any heterogeneity and repeat analysis to validate
results. However, samples composited over large intervals or different units should be
avoided.

4.2 GEOCHEMICAL STATIC TESTS

Identifying the geochemical nature of waste samples is necessary to assess their


potential for sulphide oxidation leading to acid generation. Static tests evaluate the
balance between acid generation potential (oxidation of sulphide minerals) and acid
neutralising capacity (dissolution of alkaline carbonates and other relevant minerals).
Current best practice static tests comprise:

• acid base accounting or net acid producing potential (NAPP) test;


• net acid generation (NAG) test;
• saturated paste pH and conductivity (EC); and
• total and soluble metal analysis.

These procedures are discussed below (after Miller & Jeffery 1995) using common
Australian terminology and units.

13
Photo: John Johnston
Acid leachate from waste rock stockpiles, Mt Lyell, Tasmania.

NET ACID PRODUCING POTENTIAL (NAPP)

Net acid producing potential (NAPP) is determined by subtracting the estimated acid
neutralising capacity (ANC) of a sample from the estimated total potential acidity of
the sample. Neutralising capacity comprises primarily carbonates and silicate minerals
and, to a lesser extent, exchangeable cations on clays. Potential acidity is calculated
from the content of total or reactive sulphur in the sample.

The NAPP procedure consists of three components:

• maximum potential acidity (MPA)


• acid neutralisation capacity (ANC)
• sample classification

MPA is calculated by determining the sulphur content of the sample as a percentage


and multiplying the sulphur content by a conversion factor. Assuming complete
oxidation of all the sulphur, the conversion factor generates kilograms of acid that can
theoretically be produced from one tonne of material with that sulphur content.

ANC quantifies the ability of the sample to buffer or neutralise acid produced by the
oxidation of sulphur in the material. The ANC is determined by reacting a ground
sub-sample with a known quantity of acid (usually hydrochloric acid or sulphuric
acid) to determine its neutralising capacity.

Interpretation of static test results involves subtracting ANC from the potential
acidity. The sample result is given as Net Acid Producing Potential (NAPP).

A sample with NAPP greater than zero is usually classified as potentially acid
forming while a sample with an NAPP equal to or less than zero is classified as non-

14
acid forming. A result close to zero (in the range + to - 2) is often regarded as a zone
of uncertainty, although the overall acid generation potential will be small.

While the NAPP indicates the acid generating potential, it does not quantify the
reactivity of the material. The reactivity will determine whether acid conditions are
likely to occur in the field. However, it is important that the NAPP test (acid base
accounting) only be used as a screening process.

An understanding of the nature of sulphur present in samples is essential to


interpreting the NAPP procedure correctly. It is important to differentiate between
sulphide content and total sulphur. For example, organic sulphur frequently found in
coal wastes and sulphate sulphur in highly weathered materials are not a major source
of acid.

NET ACID GENERATION (NAG) TEST

A recently developed static test used to evaluate further the acid producing potential
of a sample is the Net Acid Generation (NAG) test. This procedure estimates the net
acid potential directly. The test offers advantages in that it is simple and fast,
requiring the minimum of laboratory equipment and can provide an indication of
sulphide reactivity and available neutralising potential usually within twenty four
hours.

NAG testing comprises the addition of a strong oxidising agent such as hydrogen
peroxide to a prepared sample and the measurement of the solution pH and acidity
after the oxidation reaction is complete. Reaction kinetics can also be monitored
during the test to assess sample reactivity.

A NAG pH result greater than 4 classifies the sample as non-acid forming. A NAG
pH result less than or equal to 4 confirms that sulphide oxidation generates an excess
of acidity and classifies the material as higher risk.

The NAG and NAPP test procedures are complementary in that the NAPP provides
the theoretical maximum potential for acid generation and neutralisation to derive the
theoretical balance, while the NAG test is a direct measure of the net result of the
balance of these reactions. The NAPP and NAG procedures alone are not sufficiently
definitive tests on which to base planning and operations decisions for mine
development.

SATURATED PASTE PH AND CONDUCTIVITY

The simplest static test is the saturated paste pH and conductivity (EC) test which
gives a preliminary indication of the in situ pH of the mine waste material and the
immediate reactivity of the sulphide minerals and acid neutralising minerals present in
the sample. A representative sample of crushed (<1 mm) mine rock is saturated with
enough distilled water to create a paste and the pH and EC of the mixture is
determined after a period of equilibration (usually twelve to twenty four hours). A
sample with a pH less than 4 indicates the sample is naturally acid regardless of the
NAPP, while an EC greater than two deciSiemens per metre (dS/m) — a measure of
electrical conductivity — indicates a high level of soluble constituents.

15
TOTAL AND SOLUBLE METAL ANALYSIS

Sulphide oxidation may enhance the solubility of metal constituents in the waste,
thereby significantly increasing heavy metal mobility and total dissolved metal
concentration in acid drainage. The chemical properties of the waste determine the
mobility of the metals, while both chemical and physical factors affect their potential
for migration from the source.

Initial screening should compare metal concentration in the solids with that of the
background soils and country rocks in the area (or with the average crustal
abundance). Statistical methods are available to determine whether any enrichment is
significant.

4.3 GEOCHEMICAL KINETIC TESTS

Geochemical kinetic tests involve established site or laboratory tests to simulate


weathering and oxidation of rock and process waste samples over time under
exposure to moisture and air. The tests can provide an indication of the oxidation rate,
time periods for onset of acid generation (lag time) and the effectiveness of control
techniques which may limit the reaction rates of oxidising material. The tests also
provide data for prediction of metal release and loading in drainage and leachate from
waste materials.

Columns and humidity cells are currently the most frequently used laboratory based
kinetic test techniques. Humidity cells have largely been developed by the coal
industry for testing overburden materials. The lag period for these wastes is typically
of short duration and standard testing procedures typically run for 8-10 weeks,
although longer duration tests are being evaluated. This is inadequate for materials
with longer lag times as frequently occur at other mine types. Column tests are
considered more representative of field conditions and have the additional advantage
of allowing remediation measures to be tested and compared, although these have to
be designed carefully. Expert assistance should be sought when selecting and
interpreting kinetic tests.

Photo: Placer Pacific Limited


Kidston Gold Mine Limited. Hydrological monitoring installations in waste dump
cover trial (paddock dumping concept).

4.4 INTERPRETATION OF TEST RESULTS

The NAPP and NAG tests are useful screening tools to assess the likelihood of acid
drainage developing through accelerated oxidation and neutralisation reactions.
Despite the level of interpretation required, NAPP and NAG tests are the preferred

16
method for characterising mine rock and process wastes for both the industry and
regulators. Static testing is the first step in the systematic evaluation of materials to be
disturbed and generated by mining operations. When interpreting the results of all
predictive testing it is important to seek professional assistance and evaluate all
testwork results in light of the prevailing extremes in field conditions. State
Government Departments of Mines, Chamber of Mines or the Minerals Council of
Australia can provide guidance.

5. SULPHIDE OXIDATION MANAGEMENT/ACID


DRAINAGE CONTROL STRATEGIES
Essential control factors include:

• understanding the physical and chemical factors in sulphide oxidation;


• geochemical characterisation of waste materials for acid generating potential;
• classifying and quantifying the acid generating risk of all materials to be
disposed of throughout the mine life;
• developing appropriate mine planning and selective handling and disposal
practices for materials of different risk;
• management commitment to implementing the preferred management strategies;
• work force training to identify and manage these different materials;
• monitoring to evaluate the performance of remediation strategies;
• evaluating monitoring effectiveness; and
• revising and evaluating strategies to mitigate sulphide oxidation.

Current best practice environmental management of sulphidic mine wastes requires


assessment of risk of acid generation in mine wastes, the early characterisation and
classification of all materials, and developing strategies to minimise oxidation of these
materials. Where significant sulphide oxidation is unavoidable, the preferred strategy
is to contain oxidation products and isolate the higher risk materials. A less desirable
option is to treat the resulting acid drainage and its impacts.

Photo: John Johnston

17
Diversion drainage around waste rock dumps and clay lining of drains (where
necessary) is important to minimise the ingress of water to the dump.

When considering acid drainage management, a distinction must be drawn between


historical mines and new mining operations. Technical options for managing acid
drainage are considerably greater at new mine sites where control strategies can be
implemented as part of the mining plan. With historical sources of acid drainage,
abatement measures may be limited by economic constraints and the severity of
impacts.

The most cost effective management strategy for sulphide oxidation is integration of
oxidation and hydrological controls throughout all stages of mine operations, from
mine planning to mine closure, and working in a coordinated manner to minimise the
risk of acid drainage developing.

Basically, control strategies require the exclusion of one or more of the inputs that
result in oxidation, that is, sulphide minerals, oxygen or water. Where acid drainage
generation cannot be eliminated, it may be treated or its release regulated to a rate that
will not significantly affect the receiving environment.

Minimising acid drainage requires control of:

• sulphide oxidation and acid generation rates by regulating oxygen, water,


bacteria or other limiting factors, for example, alkalinity;
• water percolation through the material to inhibit migration or transport of
oxidation products from the source;
• control of alkalinity and acidity balance so oxidation products and other soluble
constituents are precipitated and immobilised within the material.

Reducing oxygen availability is the most effective control on oxidation rate. A low
oxygen permeability (diffusion) cover is also likely to restrict water movement into
and through the material thus reducing both oxidation rate and product transport. Best
practice environmental management requires site-specific adaptation of local
resources and an understanding of the local environment to produce the most
appropriate form of cover.

Photo: John Johnston


Mt Lyell, Queenstown, Tasmania. Clay is an ideal cover material for the
encapsulation of waste rock dumps to inhibit water and oxygen ingress. Placing a
dump near a clay resource can reduce rehabilitation costs.

Covers to achieve this can be broadly defined as water covers and soil covers. Water
covers provide the most effective control of sulphide oxidation rates. Soil covers can

18
only approach the efficiency of water covers when a proportion of the cover material
remains saturated, which reduces the diffusion rate of oxygen through the cover.
However, soil covers also offer the advantage of reducing the water flux/transport
medium through the material. An example of effective cover use is provided in Case
Study 6. A list of treatment options is at the end of this chapter in Table 1.0.

5.1 SOIL COVERS

Soil covers may comprise low sulphide content waste rock or materials specifically
borrowed to cover waste materials and are usually selected for their particular
characteristics such as compaction rates and low permeability. Such cover materials
are often clay subsoils or oxide wastes. Synthetic membranes such as geotextile
fabrics can achieve a high efficiency in water exclusion and have been used for
covering low grade ore stockpiles to prevent oxidation prior to processing.

All available cover materials should be evaluated, based on their capacity to minimise
oxygen and water availability.

Photo: John Johnston


Spreading of clay on Magazine Waste Rock Dump at Mt Lyell. Following spreading,
clay is compacted to achieve the desired permeability (1 x 10 -9 mm/s ) to inhibit
oxygen and water permeating into the dump. The dump surface is graded to a
perimeter diversion drain to control runoff and minimise erosion. A 250 mm thick
uncompacted revegetation layer is placed over the compacted layer to seal it and
prevent erosion.

Oxygen control is best achieved if the pore space of a portion of the cover remains
saturated, or within about 10% of saturation. This layer will substantially reduce the
oxygen flux through the cover, as the diffusion coefficient of oxygen through water is
at least one order of magnitude smaller than that in air. Design of a soil cover system
may require three zones that have the following properties:

• a zone, usually the base zone, with high water retention properties, which
provides the greatest barrier to oxygen diffusion ('water retention' zone);
• a zone to act as a water reservoir to ensure that some portion of the water
retention zone remains close to saturation; and
• a surface zone that protects the cover from erosion ('barrier' zone).

In many cases the surface and barrier zones provide a growth medium for vegetative
cover which also helps to minimise erosion and control percolation. In some cases a
'capillary break' zone underlies the water retention zone to limit the upward migration
of soluble constituents to the surface which may affect the vegetative cover. In

19
practice the frequently coarse nature of waste materials and their low water content
means that they act as their own 'capillary break'.

The three zone concept was applied to the rehabilitation of the Rum Jungle mine site
in Batchelor, Northern Territory. The main objective in using the layered system on
the waste rock stockpiles was to reduce the water flux. However, post-rehabilitation
monitoring has shown that, owing to the extreme rainfall variations between the
tropical wet and dry seasons, the cover system was effective in reducing the oxygen
flux only during the wet season, due to the climatic (rainfall) extremes experienced
during the wet and dry seasons in the Northern Territory.

While reducing the water flux through waste material may not significantly reduce the
oxidation rate, installation of a soil cover has the definite advantage of reducing the
quantity of drainage emanating from the base. The stored oxidation products may also
be effective in helping to reduce the overall oxidation rate of the residual sulphide
minerals.

Photo: Renison Limited


Renison Limited, Renison Tin Mine, Western Tasmania. Sub-aqueous disposal of
tailings at Dam C. Where beaches are visible, sulphidic tailings are covered with
inert tailings to control oxidisation.

5.2 WATER COVERS

When stored under water, unoxidised sulphidic mine wastes are largely chemically
unreactive. The use of water as a cover material has a similar objective to a saturated
soil cover in that it reduces the availability of one of the principal reactants —
oxygen. The maximum concentration of dissolved oxygen in natural waters is
approximately 25 000 times lower than that found in the atmosphere (due to the lower
diffusion coefficient of oxygen in water). Once the available oxygen in water is
consumed, the rate of reaction is reduced as its rate of replacement is relatively slow.
The resultant diminished availability of oxygen is the single most effective inhibitor
to sulphide oxidation.

Water covers are more readily achieved in temperate climates. Other methods to
ensure saturation may be required in drier climates, for example, establishing a
permanent wetland on tailings impoundment surfaces, or designing a complex layered
cover to trap precipitation on the tailings surface and inhibit evaporation. While the
use of natural or artificial lakes as repositories for mine wastes may compromise other
20
beneficial uses of these waterbodies, in some instances it may be the best available
option. There are example of minesites in Canada and Sweden where lake
environments are being used for tailings and waste rock disposal. Details of
rehabilitating a mine pit by converting it to an aquatic habitat and potential water
resource are provided in a paper by Sinclair and Fawcett (see Further Reading).

As a long term management strategy, subaqueous impoundment of sulphidic wastes is


attractive as lake sediments tend to be a stable environment for sulphides. In addition
to the low concentration of available oxygen, sediments have a natural tendency to
become chemically reducing due to high organic matter levels and biological activity.

Acid generating wastes, disposed of underground as backfill in mined-out workings,


can be flooded at the end of mine life, thus minimising long term sulphide oxidation
rates.

5.3 SELECTIVE HANDLING AND ISOLATION OF SULPHIDIC


WASTES

The objective of this strategy is to isolate reactive or higher risk wastes for selective
disposal either separately or within non-reactive (lower risk) materials. Geochemical
testing enables the classification and field identification of different waste types for
selective handling and disposal. Figure 9 shows how sulphidic waste can be isolated
within a waste stockpile.

Selective handling and placement of mine wastes requires integration of sulphidic


waste management practices into the mine planning schedule, along with education
and training of the workforce to facilitate operational practices in the selective
handling of higher risk materials.

The mine plan for underground operations may incorporate successive backfilling and
selective underground disposal of sulphidic wastes, including tailings return
underground as sandfill or thickened paste, thereby reducing the amount of acid
generating material requiring disposal above ground. Incorporation of cement to bind
tailings fill has the advantage of introducing an additional source of alkalinity (as well
as strength and stability) to the material.

In some cases, it may be preferable to segregate highly reactive wastes within a


separate facility to permit intensive treatment and control strategies.

21
Photo: Placer Pacific Limited
Placer Pacific Limited, Osborne Mines (approximately 130 km south of Cloncurry,
North West Queensland). Waste rock dump showing void left in centre of dump to
'encapsulate' high suphide containing waste rock.

5.4 BLENDING

Where both acid generating and acid consuming materials are present in waste,
materials blending (or mixing or co-disposal) may be a viable option, usually
conducted in conjunction with other prevention strategies such as encapsulation
and/or soil covers. The objective is to balance alkalinity and acid generating potential
to minimise the risk of net acid generation. The operating practices involved in the
mining operation will largely determine the feasibility of this option.

Photo: Placer Pacific Limited


Placer Pacific Limited, Kidston Gold Mine. Waste dump cover trial showing
'paddock' dumping concept.

Other alkaline products may also be available for blending. In the coal industry,
flyash or kiln dust may provide a useful source of alkalinity which can be
incorporated into pyritic overburden and coal waste rejects to minimise acid drainage
potential.

At base metal mines coarse lime rejects from the processing plant may contain
residual alkalinity which could be blended with acid producing materials during
disposal. Practices such as these can assist in reducing the overall risk of acid
drainage.

Figure 9.0. Schematic of isolation strategy

Construction of surface drains to divert runoff and minimise the upstream catchment
of operational areas is an effective management strategy to reduce the volume of
surface water on site which could come into contact with acid producing material or
contaminated water. This will reduce the potential for off-site environmental impacts
from acid drainage.
22
With many base metal mines pyrite removal may be an option through incorporation
of pyrite flotation cells into the mill circuit. This will significantly lower the pyritic
content of tailings and reduce the potential for sulphide oxidation while producing a
saleable product or small amount of pyrite rich material for selective disposal.

5.5 BACTERIAL INHIBITION

Certain bacteria (APB) are known to increase greatly the rate of acid production from
pyritic materials. Bactericides have been developed which inhibit the growth of these
micro-organisms. Their primary effect is minimising the catalytic role played by the
bacteria in converting ferrous iron to ferric iron under acid conditions (where ferric
iron is the principal oxidant).

Because this is only a short term solution and only partially effective, bactericides
need to be part of an integrated systems approach to managing sulphidic wastes. Most
frequently they are applied to temporary ore stockpiles or to waste rock stockpiles to
delay the onset of acid conditions, or to reduce secondary treatment costs such as lime
dosing of drainage/runoff waters, while other more permanent solutions are
implemented.

To date this technology has largely been used for rehabilitation in the US coal
industry to assist establishment of an active vegetation cover prior to the onset of
significant acid generation.

23
DESIGN
STRATEGY FEATURES
APPROACH
Proactive: 1. Selective AMD producing waste (or tailings) is
Preventative Handling/Encapsulation selectively handled and surrounded with
Measures benign, non acid producing materials
like oxide mine wastes to limit flow of
air and water into waste and AMD flow
out. Benign material is used for
foundation and containment bunds to
form a cell structure with the surface
covered with compacted benign material
(usually clay) either staged or at the
closure of the mine.
See Figures 10.0 and 11.0
2. In-pit Disposal Similar in concept to encapsulation.
Method is useful where a mined out pit
of sufficient size is available. With
effective mine planning an early closure
of one of a series of mined pits allows
for in-pit disposal of AMD wastes.
Depending on permeability, the pit wall
and floor may require encapsulation
layers, including a surface cover. In
temperate climates, surface cover may
be provided by a water cover of
sufficient depth instead of a compacted
fill material. Surface water management
and diversion options should be part of
the design features.
See Figures 10.0 and 11.0
3. Blending/Mixing/Co- Involves the blending/mixing and co-
disposal disposal of AMD wastes/tailings with
benign non acid producing materials or
even acid neutralising materials.
Historically used at coal mining
operations, co-disposal consists of
constructing small cells within a waste
dump which are rapidly filled and
covered to reduce AMD generation and
water ingress. Cells are constructed
using benign waste materials of low
permeability.
Option of mixing tends to be for mine
waste rock rather than tailings and relies
on the neutralising capacity of the mixed
benign material rather than
encapsulation. See Figures 10.0 and
11.0
Proactive: 4. Micro-encapsulation A process of coating certain mine

24
Preventative wastes to prevent pyrite oxidation.
Measures Mechanism involves leaching the waste
with a phosphate solution with hydrogen
peroxide. The surface of pyrite is
oxidised by the peroxide to release iron
oxides which react with the phosphate
solution to form a phosphate precipitate.
This forms a passive surface coating
over the waste rock fragments.
5. Uncontrolled Placement Option for marginal acid producing
with Downstream wastes where subsequent acid drainage
Collection and Treatment is recovered and treated downstream.
of Water Collection/recovery systems can include
catchment ponds, drains, trenches and
groundwater bores. Treatment/disposal
systems include chemical treatment (eg
lime dosing) controlled release and
dilution by adjacent streams,
evaporative disposal, process reuse, and
wetland filter treatment. Each system
should be considered in the context of
site specific characteristics. This option
can be applied as a possible remedial
measure. See Figure 12.0
Reactive: 1. Covering Option involves the construction of a
Remedial low permeability cover over an existing
Measures waste rock dump, mainly using locally
available borrow or benign waste, to
reduce the infiltration of surface water
and infusion of air into the dump. Cover
design should include detailed
hydrological and seepage modelling to
optimise cover thickness. See Figure
12.0
2. Downstream Collection Option as a remedial measure for
and Treatment of Water marginal acid producing wastes where
subsequent acid drainage is collected
and treated downstream. See Option 5
above for details and Figure 12.0.
3. Removal Removal of waste rock or tailings is an
option but is not usually considered in
view of the costs involved. There may
be advantages in selective removal of
severe AMD waste from an existing
dump for isolation either within an open
cut, void or at a suitably prepared waste
dump site.
Table 1.0: A Summary of Engineering Options to Manage Acid Mine Drainage for
Mine Waste Dumps and Tailings Storages (after Marszalek 1996)

25
Figure 10.0: Methods for AMD control in waste rock dumps using a range of
encapsulation and co-disposal options. (after Marszalek 1996)

26
Figure 11.0: Methods for AMD control in tailing
storages using a range of encapsulation and co-
disposal options. (after Marszalek 1996)

27
Figure 12.0: Methods for AMD control in waste rock
dumps using covers and downstream collection and
treatment of water options. (after Marszalek 1996)

6. TREATMENT STRATEGIES
Whilst management to reduce the production of acid drainage is the preferred option,
it may be impossible to stop it completely. Treatment of the acidic effluent will
normally be necessary to control off-site environmental impacts. The mining industry
has developed a number of treatment strategies to minimise the potential for these
impacts.

6.1 TREATMENT SYSTEMS

The most common strategy for treatment of acid drainage is collection and
neutralisation with alkaline reagents such as:

• limestone (calcium carbonate);


• hydrated lime (calcium hydroxide);
• quick lime (calcium oxide);
• caustic soda (sodium hydroxide); or
• caustic magnesia (magnesium oxide),

followed by a settlement stage to recover the fine metal precipitates (hydroxides). The
appropriate alkaline reagent will depend on cost and availability, and the target pH of
the final effluent.

28
Photo: John Johnston

Acid drainage is toxic to stream aquatic life where it leaches from waste rock
stockpiles and adits, Comstock Creek, Western Tasmania.

Neutralisation is effective and reliable. Drawbacks include:

• high establishment and ongoing costs (a maintenance intensive system — cost


of limestone or other alkaline reagents make treatment one of the most
expensive remediation options);
• it treats the resultant effluent instead of avoiding its production; and
• it produces a precipitate sludge that requires separate disposal.

As a result of these factors, treatment is best considered an interim measure. Ongoing


treatment of acid drainage post mine closure is not an attractive option due to the time
scales involved, typically measured in decades or centuries, thus presenting a long
term liability for the industry. Where feasible, treatment should be replaced by more
cost effective abatement measures directed at controlling the source of acid
generation.

METAL RECOVERY

When present in sufficient concentration, dissolved metals in acid waters may


represent an economic resource. Recovery of metal through solvent extraction and
electrowinning or other extraction technologies may be commercially viable. A
discussion of this technique appears in the Office of the Supervising Scientist Report
108 (listed at the end of this booklet under Further Reading). Under some
circumstances it may be economic to accelerate the oxidation of waste heaps to
provide a metal enriched effluent to improve the feasibility of metals recovery.

Solvent extraction, leaching and electrowinning technologies generally result in a


raffinate waste which may be more acid than conventional acid drainage, containing
essentially the same constituents minus the metal targeted for recovery. Conventional
alkaline treatment technologies are usually necessary to neutralise such effluent after
metal has been extracted (see also OSS Report 108).

6.2 PASSIVE SYSTEMS

Passive alkalinity producing systems are designed to introduce alkalinity into drainage
waters. These systems are best suited to relatively low flow and low acidity waste
streams but may also be used to raise alkalinity levels in waters infiltrating sulphide
wastes.

29
PASSIVE ANOXIC LIMESTONE DRAINS

In order to reduce the maintenance requirements of mechanical treatment systems,


passive alkalinity addition systems have been developed. One such treatment system
which has been applied with varying results is the Passive Anoxic Limestone Drain
(PALID). Acid drainage is usually passed through a constructed channel of coarse
limestone gravel which excludes oxygen (that is, under anaerobic conditions). Oxygen
exclusion minimises the precipitation of metal hydroxides within the drain and on the
limestone particles, which would reduce its efficiency.

Subsequent aeration and ponding of the discharge from the drain results in
precipitation of metal hydroxides in the settling pond and a clear decant (supernatant).
Passive Anoxic Limestone Drains are generally considered to have a relatively short
effective life as the alkaline materials in the drain will be consumed requiring some
ongoing maintenance of the system. Generally, they do not suddenly cease to
function; their efficiency reduces over time.

SUCCESSIVE ALKALINITY PRODUCING SYSTEMS

Successive alkalinity producing systems (SAPS) are designed to avoid some of the
problems of PALIDs by increasing alkalinity levels in 'clean' runoff/drainage streams
and introducing these streams into acid drainage streams.

The alkalinity is achieved by using either anoxic limestone drains or open limestone
drains. Problems associated with sludge and armouring decreasing the efficiency of
alkaline materials are avoided under the anoxic conditions. Conventional aerobic
settling ponds follow these systems, providing for the collection of precipitated
sludges. For greater efficiency SAPS can be followed by wetland 'polishing' systems,
which generally produce high quality (low metal concentration) effluents.

Figure 13.0 Typical passive anoxic limestone drain—


'Cal Filter'

Panelling of Cal Filter reduces flow rates, allowing for longer resident times. The
panelling specifications are customised to obtain optimal resident times, taking into

30
account the gradient and acidity of water to be treated.
(Courtesy of Limephos Enterprises Pty Ltd, Atherton, Queensland.)

Figure 14.0 Cross section through an anoxic limestone


drain

Figure 15.0 Successive alkalinity producing systems—


Mt Lyell Remediation system

Figure 16.0 Successive alkalinity producing systems—


Mt Lyell Remediation system

31
WETLAND TREATMENT SYSTEMS

Wetlands provide an alternative low maintenance passive treatment system and can
usefully be combined with other treatment systems, including PALIDs and SAPS, as a
final treatment stage for previously treated acid drainage.

Wetlands rely on a number of chemical and biological processes to attenuate acidity


and metals. Many of the physio/chemical/biotic processes which occur in wetlands are
poorly understood; however, the net effect is the removal of metals which become
complexed in the substrate, and an increase in the pH due to alkaline materials and/or
sulphate reduction. An outline on the use of wetlands for AMD is provided in Jones
D. R et. al. 1995 and Noller B. N et. al. 1996 (see Further Reading).

Photo: Renison Limited

Water sampling of Lake Pieman on the west coast of Tasmania, to assess the possible
impact of tailings dam discharges.

Wetlands can also be established on the surface of tailings impoundments to maintain


a saturated and reducing environment, thereby reducing oxygen availability to
sulphide waste materials beneath.

A further strategy to enhance wetland treatment for acid drainage (high sulphate
waters) currently being trialed experimentally is the use of sulphate reducing bacteria
(SRB). These bacteria are naturally occurring in wetlands and sediments and
experimentally have been utilised to treat acid streams.

Design criteria for wetlands are dependent on the flow rates required to be treated,
acidity, and metal concentrations in the acid drainage. Some wetland designs
incorporate a pre-treatment of effluent with alkalinity producing systems to reduce
acidity levels and enhance metal precipitation/reduction.

Limitations to wetlands establishment include difficulties in treating large flows,


physical relief and land available for large wetland ponds. Under some circumstances
the substrate may become clogged with precipitate sludges or saturated with metals
thereby reducing the wetland's performance over time.

32
In these cases the wetland can be drained, the sediments removed and the wetland re-
established. This highlights the need for ongoing management of wetlands used for
such purposes.

33
7. MONITORING STRATEGIES
Monitoring is an essential component of sulphidic waste management. Early detection
of significant sulphide oxidation is critical to the successful implementation of
avoidance and control strategies, and is integral to a best practice risk management
approach to the problem. Essential components of a monitoring program comprise:

• background studies to identify environmental values requiring protection,


including ecosystem characteristics and catchment water quality;
• classification of all materials during the mine development and operational
phases to provide information for waste management;
• point source monitoring to locate any sites of sulphide oxidation and acid
generation;
• monitoring of catchments and groundwater systems upstream and downstream
of mine operations to determine the nature and scale of any off site impacts and
enable the recognition of long term and short term trends; and
• monitoring of control/prevention strategies to determine effectiveness, which
may need to continue post mine closure.

Both flow volume and concentration load data are necessary to characterise acid
drainage adequately and to evaluate potential off-site impacts.

Photo: Renison Limited

Renison Tin Mine. Measuring water levels in piezometers on tailings Dam B to collect
data for modelling tailings dam hydrology.

The most common indicator of sulphide oxidation is sulphate (SO4). Some other
possible indicators are iron, copper, zinc, manganese, lead and cadmium, and major
ions such as calcium, magnesium, aluminium, sodium and potassium. High sulphate
concentrations in drainage and a reduction in alkalinity, even when the pH is still
about neutral, highlights sulphide oxidation. Decreases in pH and increases in metal
concentration may lag behind the increases in sulphate concentration due to buffering
and neutralisation processes.

Specific variables such as total dissolved carbon and alkalinity should be determined
for receiving waters to provide information on their buffering capacity. This
information is useful in predicting both the potential and actual impact on receiving
waters.

34
Monitoring of operations should include a detailed inventory of all waste types,
materials classification and production levels:

• static and kinetic testing data to classify wastes and identify selective handling
requirements;
• production volumes of respective waste rock types; and
• storage location and date of emplacement of identified mine wastes.

Non-acid producing rock types, and the presence of carbonates, should also be
recorded and their suitability as cover materials assessed, as they may be useful in
remediation strategies.

Waste rock stockpiles containing higher risk material require monitoring to assess
stockpile behaviour. Monitoring of oxygen and water flux conditions may be required
to evaluate the performance of the management strategy. Where rehabilitation
strategies include the use of covers, monitoring of cover performance may include:

• oxygen ingress diffusion rates determined by analysis of oxygen partial


pressures within a stockpile;
• temperature profiles through a stockpile which indicate if exothermic oxidation
reactions are occurring;
• water flux through a stockpile, which may be determined by lysimeters and/or
piezometers; and
• physical stability (cracking, erosion etc.) of any cover material.

Long term monitoring is an important consideration as many mining landforms such


as waste rock stockpiles, tailings impoundments and mining voids may become
permanent landscape features with potential for long term environmental impact.

Photo: John Johnston

Monitoring of acid leachate from waste rock stockpile for flow, conductivity and pH
prior to rehabilitation, Mt Lyell, Tasmania.

The major factors that determine the level of long term environmental risk posed by
acid drainage at closed minesites are:

• the characteristics and behaviour of the mined materials;


• the geographical characteristics of the site including its location; and
• the regional climatic conditions.

35
Therefore, all of these factors require monitoring to assess fully the level of risk. The
results will provide useful information for mine closure planning. The mine closure
plan should incorporate strategies to minimise the risk of long term impacts on the
environment from acid drainage. As far as possible, passive management systems
should be incorporated to reduce the need for ongoing maintenance and costs.

CONCLUSION

Mining is normally a temporary land use to recover resources for the benefit of the
community. This benefit may be offset substantially if poor management results in
long term degradation of the mined area and surrounding environment from, for
example, the effects of acid mine drainage.

Best practice environmental management of sulphidic mine wastes must involve a


risk management approach incorporating:

• assessing mine waste characteristics for sulphide oxidation potential;


• assessing environmental risk based on site-specific factors;
• implementing mine design principles and mine management practices to avoid
or minimise these risks;
• identifying environmental values to be protected and designing monitoring
programs for early identification of sulphide oxidation/acid drainage;
• mine planning and mine closure strategies that minimise the potential for off-site
impacts associated with sulphide oxidation and acid drainage; and
• ensuring long term effectiveness of the management strategies implemented.

Where the risk of significant impact is high, the need for subsequent high cost
remediation strategies can be reduced by building prevention measures into
environmental management systems during mine planning and development.

36
The initial step in controlling sulphide oxidation is the early characterisation of higher
risk (acid producing) materials, and materials with potential for use in remedial
measures. An inventory should be developed, updated and refined to characterise and
quantify mine wastes adequately throughout the life of the mine. The subsequent
incorporation of this information into block modelling and mine planning will enable
the development of effective strategies to manage sulphide oxidation in these wastes.

Acid base accounting provides an appropriate means of screening material to


characterise its acid producing potential. Kinetic testing is also required to estimate
the rate of oxidation, potential oxidation product loads, and metal concentrations in
any acid drainage from these materials. However, sampling protocols are critical to
ensure that the risks are quantified with a high level of confidence. Acid base
accounting should continue through all phases of project evaluation and development.

Best practice environmental management may require the implementation of a range


of site-specific measures based on local conditions and available resources to
minimise the risk of environmental impact from acid drainage. The most effective
management strategies will require a thorough understanding of the factors
contributing to the formation of acid drainage, the specific characteristics of the waste
material being dealt with, and the range of strategies available to control and/or treat
acid drainage where effective control is not achievable. More research is needed to
identify answers to the suppression and treatment of acid mine drainage, in both
generic and site-specific aspects. Consistent with the best practice approach, many
companies are expending considerable resources on research and development
associated with sulphide mine waste management and the control of acid drainage.

37
REFERENCES AND FURTHER READING
Bennett, J.W., Gibson, D.K., Harries, J.R., Pantelis, G., Ritchie, A.I.M., Verhoeven,
T.J., Alcock, J., Henkel, H. and Applegate, R.J. 1989.'Rehabilitation of the Rum
Jungle Mine Site' in Proceedings of Joint Canadian Land Reclamation
Association/American Society for Surface Mining & Reclamation Meeting.
'Reclamation—A Global Perspective', Calgary, Alberta. pp27—31.

Bennett, J.W., Gibson, D.K., Ritchie, A. I. M., Tan, Y., Borman, P. G. & Jonsson, H.
1994, 'Oxidation rates and pollution loads in drainage; correlation of measurements in
a pyritic waste rock dump', in Proceedings of the International Land Reclamation and
Mine Drainage Conference and Third Conference on the Abatement of Acidic Mine
Drainage, Pittsburgh, USA, 1994. vol. 1, pp. 400-409, USA Department of The
Interior, Bureau of Special Publications, SP O6A-94.

Bennett, J.W, and Pantelis, G., Second International Conference on Abatement of


Acidic Drainage, Montreal, 16—18 September, 1991.

British Columbia Acid Mine Drainage Task Force 1989, Draft Acid Rock Drainage
Technical Guide (Volume 1), British Columbia Acid Mine Drainage Task Force
Report.

Department of Minerals and Energy 1995, Technical Guidelines for the


Environmental Management of Exploration and Mining in Queensland, Queensland
Government 1995.

Ferguson, K. D. & Erickson, P. M. 1988, 'Pre-Mine Prediction of Acid Mine


Drainage' in Environmental Management of Solid Waste, eds W. Salomons & U.
Forstner, Springer-Verlag, Berlin.

Filipek, L., Kirk, A. & Schafer, W. 1996, 'Control technologies for ARD' in Mining
Environmental Management, December 1996, pp. 4-8.

Grundon, N. J. & Bell, L. C., (eds) Proceedings of the Second Australian Acid Mine
Drainage Workshop, Charters Towers Queensland. 28-31 March 1995, Australian
Centre for Minesite Rehabilitation Research: Brisbane, Australia.

Harries, J. 1997. Acid mine drainage in Australia: its extent anf future liability,
Supervising Scientist Report 125, Supervising Scientist, Canberra.

John Miedecke and Partners Pty Ltd, 1996. Remediation options to reduce acid
drainage from historical mining operations at Mount Lyell, Western Tasmania,
Mount Lyell Remediation Research Demonstration Program, Supervising Scientist
Report 108, Supervising Scientist, Canberra.

Jones, D. R. & Chapman, B. M. 1995, 'Wetlands to Treat AMD - Facts and Fallacies',
in Proceedings of the Second Australian Acid Mine Drainage Workshop, Charters
Towers Queensland. 28-31 March 1995, eds N. J. Grundon & L. C. Bell, pp.127-145,
Australian Centre for Minesite Rehabilitation Research, Brisbane, Australia.

Lynch, A. J., Taylor, A. & Varas, C. A. 1994, 'Solvent Extraction Boom in Latin
America' in Engineering and Mining Journal, Vol 195, December 1994, pp. 18—21.

38
Marzalek, A. S. 1996, 'Preventative and Remedial Environmental Engineering
Measures to Control Acid Mine Drainage in Australia' in Preprints of Papers,
Engineering tomorrow today. The Darwin Summit, National Engineering Conference,
Darwin, Northern Territory, 21-24 April 1996, The Institute of Engineers Australia,
Barton ACT 2600 Australia.

Mc Quade, C. V., Johnson, J. F. & Innes, S. M., Review of historical literature and
data on the sources and quality of effluent from the Mount Lyell lease site, Mount
Lyell Remediation Research Demonstration Program, Supervising Scientist Report
104, Supervising Scientist, Canberra.

Michaelsen, D. V. and Macmillan, S. 1996 'Managing acid drainage at PT Kaltim


Prima Coal, Kalimantan, Indonesia: A case study from the wet tropics.' in
Proceedings Volume 2, pp 361-371. 3rd International and the 21st Annual Minerals
Council of Australia, Environmental Workshop, Newcastle NSW. 14-18 October
1996, The Minerals Council of Australia, Dickson ACT 2602 Australia.

Miller, S. 1992, 'Acid Mine Drainage from Mine Waste and Mining Operations' in
Proceedings of the Acid Mine Drainage Workshop, Strahan, Tasmania July 27-30,
pp. 39-89, Tasmanian Chamber of Mines Ltd., Mining and Mineral Centre, 32 Davey
St Hobart.

Miller, S., Jeffery, J. 1995, 'Advances in the prediction of acid generating mine waste
materials', in Proceedings of the Second Australian Acid Mine Drainage Workshop,
Charters Towers Queensland. 28-31 March 1995, eds N. J. Grundon & L. C. Bell, pp.
33-43, Australian Centre for Minesite Rehabilitation Research, Brisbane, Australia.

Miller, S. D., Jeffery, J. J. & Wong, J. W. C. 1991. 'Use and misuse of the Acid Base
Account for 'AMD' Prediction', in Proceedings of the Second International
Conference on the Abatement of Acid Drainage, Montreal, Sept 1991.

Noller B. N. & Parker, G. 1996, 'Design of Wetland Systems at Mining Projects in the
Tropics to Control Contaminant Dispersion from Waste Water' in Preprints of
Papers, Engineering tomorrow today. The Darwin Summit, National Engineering
Conference, Darwin, Northern Territory, 21-24 April 1996, The Institute of Engineers
Australia, Barton ACT 2600 Australia.

Orr, M. 1995, 'Development of a Waste Dump Management Strategy at Mt Leyshon


Gold Mine', in Proceedings of the Second Australian Acid Mine Drainage Workshop,
Charters Towers Queensland. 28-31 March 1995, eds N. J. Grundon & L. C. Bell, pp.
91-100, Australian Centre for Minesite Rehabilitation Research, Brisbane, Qld
Australia.

Richards, R.J., Applegate, R.J., & Ritchie, A.I.M. 1996. Chapter 14: 'The Rum Jungle
Rehabilitation Project' in: Environmental Management in the Australian Minerals and
Energies Industries in Principles and Practices. David Mulligan (ed). University of
New South Wales Press.

Ritchie, A. I. M. 1992, 'Processes which Govern Pollution Levels Generated by


Oxidation in Pyritic Material' in Proceedings of the Acid Mine Drainage Workshop,
Strahan, Tasmania July 27-30, pp. 22-38, Tasmanian Chamber of Mines Ltd., Mining
and Mineral Centre, 32 Davey St, Hobart.

39
Robertson, J. D. & Ferguson, K. D. 1995, 'Predicting Acid Rock Drainage', in Mining
Environmental Management, December 1995, pp. 4-8.

Sinclair, G. & Fawcett, M. 1994, 'Development of an Aquatic Habitat and Water


Resource in Closing Out the Enterprise Pit' in Technical Program Proceedings 1994
AusIMM Annual Conference, ed C. P. Hallenstein, Australian Institute of Mining and
Metallurgy, Calton South.

Skousen, J. G. & Ziemkiewicz, P. F. (eds) 1995, Acid Mine Drainage Control and
Treatment, West Virginia University and the National Mine Land Reclamation
Centre, Morgantown, West Virginia.

Tasmanian Chamber of Mines Ltd. 1992, Acid Mine Drainage Workshop, Strahan,
Tasmania, July 27-30, Tasmanian Chamber of Mines Ltd., Mining and Mineral
Centre, 32 Davey St Hobart

Treloar, J. H. & Wilmshurst, R. E. 1993, 'Copper Recovery by in situ Leaching,


Solvent Extraction, and Electrowinning by Gunpowder Copper Ltd, Gunpowder, Qld'
in The Sir Maurice Mawby Memorial Volume, Volume 1, eds J. T. Woodcock & J. K.
Hamilton Monograph No 19, pp 724-729, Australasian Institute of Mining &
Metallurgy, Carlton South, Vic Australia.

United States Environmental Protection Agency 1994, Technical Document Acid


Mine Drainage Prediction, Washington DC 20640.

United States Department of the Interior 1994, International Land Reclamation and
Mine Drainage Conference and Third International Conference on the Abatement of
Acidic Drainage, vol. 1-4, Pittsburgh PA, April 24-29, 1994. Bureau of Mines Special
Publication SP 06a-94.

White, L. D., Meka, Z., MacDonald, C. A. & Miller, G. 1993, 'Solvent Extraction and
Electrowinning Capital' in Cost Estimation Handbook for the Australian Mining
Industry eds M. Noakes & T. Lanz, pp 203-216, Australasian Institute of Mining &
Metallurgy, Carlton South, Vic Australia.

40
CASE STUDY 1
GREGORY OPEN CUT COAL MINE, BHP
AUSTRALIA COAL
The Gregory Coal Mine is an open cut mine located on a lease of 6 500 ha some 60
km north east of Emerald in the Bowen Basin, Central Queensland. The strata just
above the main coal seam are pyritic. The pyrite generates acid unless preventative
measures are taken as part of the mine operations.

The climate of the area is sub-tropical, semi-arid with about 80% of the annual
average rainfall of 620mm falling in the warmer months of October to May. Annual
average evaporation exceeds rainfall by a factor of three.

Photo: BHP Australia Coal


BHP Australia Coal, Gregory Coal Mine, 60 km north east of Emerald, Queensland.
Rehabilitation of mine spoil in foreground with strip mining in background.

Assessment of the exploration drilling for the Gregory Mine in the late 1970s
identified the potential for acid generation. Therefore, mine planning and more
particularly operational planning identified measures to minimise the generation of
acid. Operational measures were introduced from day one of mining.

MEASURES TO REDUCE ACID GENERATION

The following outlines the operational measures introduced to reduce acid generation.

• Temporary Reduction of Acid Generation from Wastes by Covering with


Water

The fine wastes and tailings from the coal preparation process are slurry pumped to
tailings impoundments or to mined out pits. The disposal of these wastes by covering
with water reduces the rate of acid generation.

The supernatant water from these disposal areas is returned to the preparation plant
for reuse in coal washing. Although this returned water usually requires some
neutralisation in a lime treatment plant, the deep sub-aqueous disposal of the tailings
reduces the total volume of acidic water generated at the mine.

Reuse of the water in the tailings disposal cycle also reduces the total demand for raw
water on local water resources by the mine.

• Permanent Covering of Wastes with Benign Material.


41
All coal-like wastes and particularly the coarse rejects from the coal preparation plant
are returned to mine pits. This negates the construction of reject/waste dumps with a
very high potential for acid generation. The dumped rejects are subsequently covered
with spoil derived from the excavation of the next mine pit-strip. The rejects are
returned to mine pits as part of the coal haulage movements, generally maintaining a
shorter haulage route and quicker haulage cycle.

Photo: BHP Australia Coal


BHP Australia Coal, Gregory Coal Mine, near Emerald, Queensland. Rejects are
returned to mine pits which removes the need to build reject/waste dumps. These
wastes are subsequently covered with spoil.

As no reject dumps are constructed, there is no requirement for special rehabilitation


of reject dumps. This significantly reduces overall rehabilitation costs and minimises
the future risk of acid generation from massive reject dumps.

• Selective Handling

Selective handling and some rehandling of pyritic spoil materials is necessary to


reduce acid generation in the mine spoil. Selective handling was identified prior to the
start of mining, this practice is a usual practice at the Gregory Mine and not a special
practice with additional costs.

The aim in the selective handling is to place the pyritic spoil deep in the spoil profile.
This is generally accomplished by rehandling by the dragline of identified pyritic
spoil into the first pass spoil valleys. The rehandling places the pyritic spoil at depth
and reduces the overall regrading requirement of the spoil as part of the mine
rehabilitation activities. Deep placement of the potentially acid generating spoil also
ensures more successful revegetation of the mine spoil because benign non-toxic
materials are used as a cover.

Photo: BHP Australia Coal


BHP Australia Coal, Gregory Coal Mine, near Emerald, Queensland. With the use of
benign materials as a cover, and deep placement of acid generating spoil, this allows

42
more sucess with revegetation of mine spoil. Weak vegetation establishment due to
poor placement of spoil materials will require further patch-up works.

Should acid generating spoil remain on the surface, or be exposed during normal
regrading of spoil, the regrading of the spoil materials is altered in the field to ensure
burial of the pyritic materials. Small equipment has also been used to "patch-up"
rehabilitation where acid generation has caused problems or where topsoil cover is
insufficient.

WATER MANAGEMENT STRATEGY

Recognition of the potential for acid drainage at Gregory Mine during mine planning
also led to the development of a sound, comprehensive water management strategy.
Drainage unaffected by mining activities is diverted away from or directly through the
mine area. This drainage is therefore not contaminated by the mine workings.

Mine drainage, and particularly mine process water and water which accumulates in
mine pits, potentially is contaminated by acid mine drainage. Therefore, this drainage
is captured and reticulated back to the mine industrial area where it is available for
dust suppression and if required for use in the coal preparation plant.

Figure 1.0 Schematic water management system at


Gregory Open Cut Mine.

This water management system reduces the requirements for raw water from external
supply sources and reduces the need to release water from the mine area into local
water resources. Therefore, downstream water users are not put at risk; releases of
mine water occurring only when there are flood flows in local streams and mine
waters are well diluted.

The early identification of the potential for acid generation from near-coal strata at
Gregory Mine has allowed the development and implementation of measures to
mitigate the risks of acid formation. The implementation of these plans has allowed
mining costs to be controlled, site water consumption to be moderated and site
rehabilitation costs to be reduced.

43
CASE STUDY 2
CADIA - WASTE BLOCK MODELLING FOR ACID MINE DRAINAGE
PREDICTION AT THE CADIA PROJECT

INTRODUCTION

Newcrest Mining Limited has utilised a waste block modelling approach (based upon
protocols developed by AGC-Woodward Clyde) at its Cadia project in New South
Wales to identify potential acid generating material within the deposit.

The Cadia gold/copper deposit is located 25 km south of Orange, NSW. Average


annual rainfall is 800 mm which is spread evenly throughout the year. Average annual
evaporation is approximately 1 100 mm. The deposit contains approximately 200 Mt
of ore and 270 Mt of waste rock that will be mined over a 12 year period.

The area is historically an active mining area with exploration and mining dating back
more than 140 years to the discovery of gold and copper in 1857.

Photo: Newcrest Mining

Aerial view of the Cadia Project.

Photo: Newcrest Mining

Cadia Hill viewed from adjacent hills.

44
Geology

The Cadia Hill mineralisation is indicative of a relatively oxidised and low sulphide
system, reflected by low abundance of pyrite, typically less than 1%. The sulphide
minerals occur in two associations — as inclusions in quartz veins and as fine to
medium ground disseminations in the monzonite porphyry and volcanic host rocks.

The dominant sulphide mineralogy is pyrite (FeS2), chalcopyrite (CuFeS4) and minor
bornite (Cu5FeS4). The sulphides are generally fracture controlled and associated with
sheeted quartz veins. Where carbonate is present in the waste rock it is predominantly
calcite, manifesting as cross-cutting veins and as a pervasive alteration of the waste
rock minerals.

Sampling strategy and testwork

Based on early identification/classification of the mine waste, a mine waste schedule


has been formulated to facilitate selective placement of waste rock within the waste
stockpile and so minimise potential acid drainage.

An extensive geological database on the site has been compiled from logging of
drillcore at one metre intervals. The general sampling approach adopted is shown in
Figure 2.0.

45
Figure 2.0 Evaluating Acid Mine Drainage Potential

46
Criteria for the selecting material from the database for testing include:
• lithology;
• sulphide abundance;
• sulphide form;
• carbonate abundance;
• carbonate form;
• oxidation state;
• alteration mineralogy; and
• structure.

Each of these criteria has the potential to influence the acid generating or neutralising
capacity of individual rock types.

Thirty six separate combinations of physical and mineralogical characteristics for the
eight waste rock types were identified which had the capacity to influence the acid
drainage potential.

On the basis of results from static testing, four waste types were selected for longer
term kinetic testwork:

• monzonite high pyrite and low carbonate


high carbonate and low pyrite
• volcanic high pyrite and low carbonate
high carbonate and low pyrite

Photo: Newcrest Mining

Long term column testwork has been initiated to further assess the quality of waste
rock leachate.

47
Photo: Newcrest Mining

Newcrest Mining, Cadia project near Orange, New South Wales. Drill core used in
the environmental/ geochemical waste characterisation study.

Classification of waste rock

The static testwork identified three main waste types:

Type I potentially acid forming (total sulphur >0.25% and MPA:ANC ratio of
1:1 or less);
Type II indeterminate (total sulphur >0.25% and MPA:ANC ratio of 1:3 or
less); and
Type III non-acid forming (total sulphur <0.25% and MPA:ANC ratio of 1:3 or
greater).

Mine Block Modelling

These waste rock sample characteristics provide a set of parameters for waste
classification into potentially acid producing, indeterminate or acid neutralising. Each
waste rock type is classified in these terms and the information integrated into the
database.

Once classified, the waste rock distribution and volume as a proportion of the total
waste volume is determined by mine block modelling and percentage volumes are
estimated for the three 'types' (Type I-III) of waste.

With the volume and position of acid producing material defined, this information is
integrated into the mining schedule to determine how much and when potential acid
drainage material will be produced during mining.

48
Results

Figure 3.0 shows the distribution of the waste block types with respect to the ore.

The static testwork and block modelling has identified a small percentage of
potentially acid producing waste than can be located in discrete blocks within the
proposed pit area. Using the mining schedule, this material can be identified during
mining and selectively handled for placement in specific cells in the waste stockpile
for blending with waste rock of high neutralising potential.

A significant volume of non-acid forming and acid neutralising waste can also be
blocked out and selectively used in the construction of the waste stockpile, for
example, to encapsulate and/or blend with potentially acid forming waste, and for use
as rock armour on the outer batters of the waste stockpile.

Conclusions

The use of a sampling program based on geological characteristics to test for acid
producing rock (and block modelling to identify the spatial distribution and amount of
acid producing material at Cadia) has resulted in:

• minimising the amount of waste material to be selectively handled;


• developing cost effective waste handling programs; and
• integrating the strategy for handling acid-forming waste with the operational
mine schedule.

Figure 3.0 Typical Cross-Sections ANC/MPA Block Model

49
CASE STUDY 3
BHP MINERALS, CANNINGTON

Location

North-west Queensland approximately 200 km south east of Mt Isa.

Climate

The climate is sub-tropical, characterised by a relatively short and unreliable wet


season. Average annual rainfall is 440 mm, falling mainly between December and
March. Average pan evaporation is 3 000 mm per annum.

Average daily air temperature ranges from 25°-35°C in December to10°-25°C in July.

Metals Produced

Silver, lead and zinc as sulphide concentrates.

Type of Operation and History

Cannington silver was discovered in 1990 by drill testing of a magnetic anomaly


identified during a regional aeromagnetic survey. A decision was taken to proceed
with the development of the deposit in 1996 following positive results from a
feasibility study which included an underground bulk sampling program undertaken
in 1994. Construction of permanent surface facilities commenced in 1996 and
commissioning is scheduled for November 1997. The mine will produce
approximately 1.5 million tonnes of ore per annum over a mine life of approximately
20 years.

Sources of acid drainage

The project will produce both waste rock and mill tailings. A proportion of the waste
rock and the tailings material will contain reactive sulphide minerals in concentrations
that may produce acid conditions if not managed appropriately. Waste rock will be
stored on the surface in engineered waste rock stockpiles. The coarse fraction of the
tailings material will be returned underground as mine backfill and only the fine
fraction will be placed in the tailings dam.

Planning to Minimise Environmental Impact

Geochemical investigations were undertaken during the exploration phase to


characterise the waste materials. Soils and subsoils to be moved in the construction of
surface facilities were also evaluated in the program. Materials were characterised
based on their Net Acid Producing Potential (NAPP), determined from the difference
between the Maximum Acidity Potential (MAP) and the Acid Neutralising Capacity
(ANC). The results were correlated with Net Acid Generation (NAG) test results.
Salinity was included as an additional parameter in the planning process.

The NAG test proved to be a cost effective measure of the acid forming potential of
sulphidic waste materials under field conditions.

50
Measures to Reduce Environmental Impacts

Five main material types were identified through the characterisation process. These
materials are listed below along with the specific management prescriptions for each
material.

Material Geochemical Management Prescriptions


Type Characteristics
1A Non-acid forming Suitable for general construction use and general fill.
Nil/low/moderate No specific geochemical constraints. Suitable for
salinity rehabilitation works
(NAG pH >4 & EC
1:5 < 0.8 dS/m)
(NAG pH >4 & EC
1:2 <1.5 dS/m)
1B Non-acid forming Suitable for general fill. Undesirable for rehabilitation
High salinity use due to salinity. Avoid placing within 300mm of
(NAG pH > 4 & EC final surfaces.
1:5 0.8-1.3 dS/m)
(NAG pH > 4 & EC
1:2 1.5-2.5 dS/m)
1C Non-acid forming Can be used for general fill provided it is isolated
Extreme salinity within the core of any embankment (for example,
(NAG pH > 4 & EC water storage). Do not place within 500mm of
1:5 > 1.3 dS/m) rehabilitation surface.
(NAG pH > 4 & EC
1:2 > 2.5 dS/m)
II Potentially acid Not suitable for construction use or general fill unless
forming placed and compacted within the core of
Low risk embankments and isolated from leaching. Do not
(3 < NAG pH < 4) place within 1m of final surfaces or outer edge of
stockpile#.
III Potentially acid Should be buried and isolated from leaching. Place
forming and compact Type III materials in layers. Locate
High risk material towards the centre of the stockpile area. Do
(NAG pH < 3) not place within 1m of final surface or within 5m of
the outer edge of the stockpile. Place compacted Type
1C over the Type III material before placing soil
cover for rehabilitation#.

EC Electrical Conductivity of a one part soil to 5 parts water slurry (1:5), or a


one part soil to 2 parts water slurry (1:2).
NAG Net Acid Generation test result
# Type II and Type III materials can be converted to Type I material by
blending with limestone or other acid neutralising materials.

51
Materials produced during the development of underground and surface works
associated with the bulk sampling program have been handled in accordance with the
above guidelines. Type III material has been placed in a stockpile adjacent to the
decline. The area was initially stripped of topsoil (200 mm) and subsoil (300 mm) and
the subgrade compacted using a self propelled vibrating roller prior to the placement
of the waste rock. The waste rock was placed and compacted in 500 mm layers until
the design height of the stockpile was achieved. The compacted material was then
covered with 500 mm of clay to reduce the ingress of rain water and oxygen. The clay
was then covered with a protective layer (1m) of Type I material.

Waste rock dump covered in clay.

Photos: BHP Minerals

Another view of clay capped waste rock at Cannington.

Test work has demonstrated that the tailings material will have a moderate sulphur
content ranging from 1 to 2.5% sulphur which will generally be balanced by the
medium to high acid neutralising capacity of the material. The results also indicate
that the reactive sulphide content is sufficiently high to cause salting problems under
exposed evaporative drying conditions. A cover will therefore be required to avoid
salt pan development and generation of dust.

Based on the work undertaken to date it is believed that a simple single layer cover
will be adequate for the majority of the tailings material. The cover will be designed
to provide adequate depth for plant growth and prevent salt migration to the surface.
Further test work is planned to optimise the design of the cover for the tailings
impoundment. An evaluation of salt profiles in the natural soils of the area indicates
that a 600 mm cover constructed from natural soils in the surrounding area will be
sufficient to achieve the rehabilitation objectives.

52
Photo: BHP Minerals

Decline stage I. Waste rock stockpile, built on limestone pad.

Photo: BHP Minerals

Type III stockpile and clay seal plus Type I material (rock) on top.

Monitoring

The geochemical stability of the existing waste rock stockpile is being monitored
utilising thermistors and oxygen probes inserted within the Type III material to
determine the effectiveness of the design. Any increase in temperature and associated
reduction in oxygen levels would be indicative of active oxidation of sulphide
minerals and the associated potential for acid generation. Runoff and leachate will be
monitored if it occurs. Since establishment of the monitoring program in May 1995
temperatures and oxygen levels have remained stable. The waste rock stockpile has
not generated any leachates to date.

A detailed geochemical monitoring program will be implemented following


commissioning of the milling operations to provide information necessary to finalise
the design of the final cover for the tailings impoundment. The tailings facility will be
a zero discharge operation.

Benefits of Management Program

Failure to adequately identify and manage the potential for sulphidic oxidation and
acid generation can result in significant impacts on the environment and substantial
clean up costs for the mine operators. The geochemical characterisation of waste
materials and the development of handling options during the planning phase of the
Cannington project will ensure that the potential for acid generation is minimised and
controlled in a cost effective manner. Environmental values will be protected during
the operation of the mine and final decommissioning will be facilitated without the
requirement for expensive post-mine remediation programs.

53
Photo: BHP Minerals

Laying thermistors and oxygen probes to monitor temperature and oxygen inside
dump.

Photo: BHP Minerals

Oxygen and temperature monitoring control box on probes within Type III stockpile.

CASE STUDY 4
PLACER PACIFIC LIMITED

A Risk Management Approach to the Acid Drainage Problem

The Placer Dome Group (PDG) is a precious and base metal mining group of
companies based in Canada, with major operating subsidiaries in Australia, Canada,
South America and the USA. In Australia, the operating subsidiary, Placer Pacific
Limited, manages five operating mines and exploration activities within Australia and
the Asia-Pacific region.

The PDG Boards of Directors have approved an environmental policy and associated
strategic plans that require acid drainage risks to be addressed fully for all operations.
This approach also requires appropriate controls to be implemented which minimise
potential acid drainage and associated environmental impacts during operations and
after closure. In accordance with good environmental management, PDG has initiated
and conducted, since 1990, comprehensive acid drainage reviews of each major
operation and exploration project. These reviews provide information to enhance
internal environmental operating plans and rehabilitation plans for all minesites.
Research projects have been established at several sites where additional study
programs are required to enable a more comprehensive assessment.

54
PDG Corporate Philosophy

The PDG Environmental Management Group has grouped its activities into four
general categories for acid drainage potential: future mines (advanced exploration
projects/acquisitions); newly designed/constructed mines; existing/operating mines;
and decommissioned/rehabilitated mines. While there are different philosophies and
strategies developed for each group, there is a set of underlying principles which
apply generically to acid drainage prediction as outlined below:

• Each minesite is unique with respect to geology, climate and site configuration
and, hence, the application of rigorous routine prescriptive practices is not
appropriate.
• Sulphide mineral reactivity is highly variable and acid drainage may not occur
at all, or be delayed for several years following material exposure from
mining. Consequently, it is never too late to assess the acid drainage potential
at existing mines using predictive methods with additional integration of
empirical site information.
• Prediction and prevention of acid drainage is the most desired and simplest
control approach for all sites.
• The results of the predictive assessments may justify changes in mining
operations and waste disposal practices to reduce the potential of acid
drainage.
• The application of predictive methods at an early stage of an advanced
exploration project will allow a phased approach to further studies and, if
necessary, detailed planning of waste rock and tailings disposal.
• Internal training of PDG technical staff in the prediction, assessment, and
control methods for acid drainage is an integral component of good
environmental management.
• Ongoing applied research is needed at sites with significant acid drainage
potential to track the success of the implemented prevention strategies and to
augment the knowledge base for the development of improved techniques
within PDG and the minerals industry.
• Continued and significant support is required for generic research through
industry task forces and at individual PDG sites to allow a better
understanding of the processes associated with acid drainage.

Risk Management Approach

The PDG program consists of a team approach composed of several key players and
facilities. These personnel vary with the classification of the project, but the common
denominator is a joint effort between corporate personnel, specialised experts and the
exploration/design/operating personnel who will eventually be required to apply the
end product. An experienced site geologist and a mine design engineer assist the
environmental personnel in the initial overview risk assessment and provide
continuity for ongoing risk assessment and control strategy development.

55
Figure 4.0 Placer approach to ARD assessments of new and existing mines

CASE STUDY 5
RENISON TIN MINE, TASMANIA
Location
Western Tasmania, in the Pieman River catchment, 135 km south of Burnie.
Climate
Wet, cool with occasional snow. Average annual rainfall of 2 252 mm spread
throughout the year. Average pan evaporation is 500 mm.
Mean monthly temperature ranges from 3 to 22°C with an annual mean of about
15°C.

56
Metals Mined
Tin
Type of operation and history
Tin as cassiterite was discovered in 1890. Various small companies operated in the
field until the current operations began in 1966. Today Renison is one of the world's
largest underground tin mines. Ore is primarily mined by flatback cut-and-fill and
benching methods using trackless diesel equipment and decline access to working
areas. A shaft was recently commissioned to access deep ores. A complex mineral
processing plant produces a cassiterite concentrate for export.
Sources of acid drainage
The tin ore contains about 45% of the very reactive iron sulphide mineral pyrrhotite,
resulting in a Net Acid Producing Potential (NAPP) of greater than 600 kg H2SO4 per
tonne.
Acid drainage is found in:
• mine waters pumped to the surface during mine dewatering (30%);
• decant and seepage waters from tailings impoundments (55%); and
• runoff from historic workings and miscellaneous sources from the current
operations, such as the ore stockpile and small waste rock stockpiles (15%).
Strategic planning to minimise environmental impact of acid drainage
generation
One of the primary considerations is to develop management strategies that will not
compromise water quality within the Pieman River catchment after the life of the
mine.
With precipitation exceeding evaporation by about 1 500 mm per annum, water
discharge from the lease is inevitable.
Treatment of acidic waters with slaked lime is a part of the mine's current practice.
Modelling of the oxidation process occurring in the tailings impoundments indicates
that, at current oxidation rates, acid waters would be discharged from the tailings
impoundments for many decades after mine closure. It would not be realistic to
maintain water treatment over this time scale.
Under these circumstances the management focus is on eliminating oxidation at
source, or reducing oxidation rates, such that water quality within the catchment will
not compromise the long term ecological integrity of the State Reserve on the Pieman
catchment downstream of the mining operation.
Monitoring
• Weekly sampling of all point source discharges from the mine (six stations);
• bimonthly sampling of ambient waters in and around Lake Pieman (nine
stations); and
• bimonthly sampling of catchment waters as part of a joint program with State
Department of Environment and Land Management, State Hydro-electric
Commission and various mines operating in the catchment (fifteen stations).

57
Map of Renison Tin Mine showing tailings impoundments and monitoring stations.

Tailings Impoundment Close-Out Model


The tailings produced during mineral processing are a composite of three separate
waste streams. One of these, Cassiterite Flotation Tailings (CFT), is low in sulphur
with a net acid neutralising capacity. It makes up 25% of the tailings by mass and has
potential as a suitable cover material on the other tailings materials.
Research has shown that CFT would quite readily support plant growth and, if
maintained at 80% saturation, would provide a very effective barrier against oxygen
diffusion by reducing oxidation rates to negligible levels.
Renison's tailings impoundment management plan is for subaqueous disposal of
tailings into dam C during mining, leaving a stable permanent pond at mine closure.
Decant waters will discharge into tailings impoundments A and then B. These dams
will be covered with CFT to a depth of 0.5 to 1 metre and wetlands developed on top.
The wetlands will be designed to act as polishing ponds for Dam C effluent before it
is discharged to receiving waters. Specific cells for the sequential bio-removal of iron
and manganese need to be developed.

Renison Tin Mine. Conceptual Model of close-out plan for


Renison tailings impoundments.

58
Research work is continuing to further develop the close-out plan. Tailings
impoundment hydrology, dam C water chemistry and wetland cell design are being
modelled for subsequent implementation. Complementing this research is a joint
mining industry funded research program on the speciation and fate of metals in the
waters of Lake Pieman.

Research on the close-out model is primarily aimed at ensuring that the community is
not left with an environmental legacy that may compromise water quality for future
generations. Knowledge gained is being used by Renison and others to improve
tailings dam management from an environmental perspective. Contemporary practice
has seen an 80% reduction in mass discharges from tailings dams, setting new
benchmarks for the future.

Renison Limited, Renison Tin Mine. Natural colonisation of wetland species on inert
tailing cover material (CFT) on dam A within two years of placement.

Photo: Renison Limited

CASE STUDY 6
MT LEYSHON

Introduction

The Mt Leyshon Gold Mine is located 24 km south of Charters Towers, Queensland.


It is a large open-pit mine, producing gold and silver. The operation is owned by Mt
Leyshon Gold Mines Limited, which is a member of the Normandy Mining Group.

Operations began in 1987 and were initially restricted to the heap-leaching of oxide
ore. Ongoing testwork revealed a much larger resource of unoxidised ore at depth,
which lead to the construction and operation of a conventional carbon-in-pulp (CIP)
plant. Throughput is currently 5.5 Mt year.
59
Climate and Geology

Mean annual rainfall at Charters Towers is 660 mm, with most rain occurring from
November to March. Most rainfall is received in high intensity storms with high
runoff rates. Average pan evaporation is 1 965 mm. Average maximum and minimum
temperatures are 30°C and 177deg;C respectively.

The Mt Leyshon orebody is hosted in an intrusive breccia and igneous complex


approximately 1.6 km in diameter. Polymetallic mineralisation (Au-Fe-Cu-Zn-Pb-Ag-
Mo-Bi) is hosted in breccias and associated intrusives. A second host, the Mine
Porphyry, accounts for approximately 15% of the mineralisation.

Development of a Rehabilitation Strategy

As found at many metalliferous mines, the primary (unoxidised) waste rock contains
varying levels of sulphides, mainly in the form of pyrite. The rehabilitation strategy
for the waste dumps must therefore be designed to minimise the potential for AMD.
This case study describes the development of a rehabilitation strategy to meet this
objective. A more complete description is given in Orr (1995).

The rehabilitation strategy for the dumps at Mt Leyshon is the construction of a


sealing layer over the bulk waste to:

1. Reduce infiltration to very low levels;


2. Prevent convective transport to oxygen; and
3. Have the potential to reduce diffusive transport of oxygen.

This strategy is shown schematically in Fig. 5.0.

Fig. 5.0. The chosen waste dump strategy for Mt Leyshon Gold Mine.

Compaction trials

A key issue in implementing this strategy was the availability of material suitable for
seal construction. The clay reserves on-site are limited and the costs of using clay
would be relatively high. Oxide heap-leach material has been used to seal some areas

60
of the dump, but the saturated hydraulic conductivities achievable through
compaction of this material are limited to about 10-6 m s-1. Compaction of in situ
waste dump material to form a low permeability seal was seen as having potential
benefits and trials were conducted on-site. Fig. 6 shows a plan of the first compaction
trial using mine waste. The four major waste types at the mine were laid in strips to a
depth of about 1 m. Three different compaction methods were run at right angles to
these strips, namely:

1. A control strip with no compaction except for that produced by the bulldozer
when pushing out the strip;
2. A fully loaded haul truck with a total weight of about 360 t; and
3. A BH-1300 impact roller (photo). When towed at a speed of around 12 km per
hour, the large steel (8t) drum of this machine impacts on the ground with great
force. The claimed advantages of this device over the more usual vibrating drum
roller are good compaction at depth and relative insensitivity to the moisture
content of the material.

Fig. 6.0. Layout of the waste rock compaction trial at Mt Leyshon Gold Mine

The BH-1300 impact roller at work compacting porphyry on a waste dump at Mt


Leyshon Gold Mine.

Testwork conducted during the trial included compaction tests (sand replacement
method) and hydraulic conductivity. Hydraulic conductivity was measured directly
using an 'infiltrometer' built on-site which measures infiltration rate over a relatively
large area (1 m2).

61
The key finding from the first compaction trial was that the porphyry waste was soft
enough to be readily compacted (none of the other waste types were shown to be
suitable for use as a seal material). Both the haul truck and the impact roller were able
to obtain high densities at depth . The use
of a truck was discontinued due to the good apparent performance of the impact roller
as compaction. Routine truck travel was considered unrealistic if a consistent
specification was required.

Modelling of long-term performance

To quantify the performance of the possible porphyry seal configurations, infiltration


was modelled using the 'SWIM' model as described in Crees et al. (1994). The role of
mathematical modelling in the assessment of rehabilitation strategies was seen as very
important, because of the cost and time required for field trials. (The very variable
rainfall received on-site means that meaningful estimation of long-term infiltration
rates via direct measurement would require several years of monitoring at least). The
results of the modelling indicate that very low levels of infiltration can be achieved
using the impact roller on the porphyry, e.g. 1.7% of incident rainfall for a 1 m thick
layer compacted to give 1x10-8 m/s saturated hydraulic conductivity.

Geochemical assessment of the seal

Having examined the physical characteristics of the porphyry material, it was


necessary to assess its geochemical characteristics. Representative samples were
collected from drill core and areas of compacted seal, and analysed for sulphur and
net acid producing potential (NAPP). The results showed that, unlike the other main
primary wastes, the porphyry has significant acid neutralising capacity.

Material Availability and Scheduling

So far, field compaction trials, infiltration modelling and waste characterisation


confirmed the feasibility of using porphyry as a waste dump sealing material. It was
then necessary to consider the availability of rehabilitation material over the life of the
mine, with respect to total availability and timing of material from the pit.

Following examination of the rehabilitation materials balance for the mine, it


confirmed availability of material to support future rehabilitation strategies for the
different sites.

Attention was next given to materials scheduling. The production of porphyry waste
from the pit is controlled by the mining schedule. Stockpiling and rehandling must
therefore be carried out to provide a supply for rehabilitation purposes as required and
rehandling can be minimised with careful planning.

Additional Construction Details

Selective Handling of Bulk Waste

Before application of the porphyry seal, oxide waste and low-sulphur primary waste is
placed around the periphery of the dump to provide additional isolation of the more
reactive material (Figure 7). The porphyry is dumped over this surface before shaping
and compaction waste porphyry for capping is selected on a blast by blast basis using
Net Acid Generation (NAG) test criteria.
62
Compaction on Dump Batters

Testwork showed that the impact roller did not achieve good compaction running
down the batters, as it tended to push the porphyry down with it. In addition, the
average 1:4 slope was too steep to allow safe compaction along the contour. These
difficulties were avoided by adopting a terraced landform (Figure 8) which reduces
the effective slope angle from 1:4 to 1:5, allowing the impact roller to travel along the
slope at optimum speed. Each compacted bench overlaps the preceding bench,
resulting in a zone of continuous compaction from the base to the top of the dump.
The face of each porphyry bench is rock armoured to minimise erosion, and the lip of
each bench is shaped to provide a graded drain. This terraced landform reduces the
potential for erosion and allows easier access for seed and fertiliser application and
maintenance.

Quality Control

Following the implementation of the waste dump sealing strategy, a quality control
program was implemented. This allows a rapid assessment of quality of work, so that
practices can be modified in response to changes in results. Also, it provides an
ongoing record of results to allow verification that commitments made in the
Environmental Management Overview Strategy (the key approval document in
Queensland) are being fulfilled. All rehabilitated areas are sampled on a 50 m x 50 m
grid. At each point, the following is recorded:
1. Seal thickness;
2. Seal surface compaction;
3. Seal sulphur content and net acid generation (NAG); and
4. Topsoil thickness
The data is recorded with survey coordinates, allowing the production at any time of a
plan of the dump with all these parameters shown. The performance of the seal is also
being measured directly using lysimeters which measure the net infiltration through
the seal.
The benefits of apply the rehabilitation strategy have been:
• approximate halving of cost per unit area of rehabilitating waste dumps;
• avoiding the need of using other rehabilitation materials that are of limited
supply;
• achieving a high quality outcome for the sealing of the dumps;
• aid the reduction of the security deposit which is based on demonstrated
performance and the cost of rehabilitation at the time of mine closure.

63
Fig. 7.0. Formation of dump batters

Fig. 8.0. Detail of compaction on dump batter

CASE STUDY 7
CLARENCE COLLIERY PTY LIMITED

Clarence Colliery is a 2.6 million tonnes per year thermal coal mining operation near
Lithgow, New South Wales. The mine is also located in the headwaters of the
Woolangambe River, a major drainage channel through the Blue Mountains National
Park.

The coal preparation plant washes coal to achieve customer ash specifications with
some 160 000 tonnes of rejects produced per annum. This reject material has a
sulphur content of 0.16% and if not well managed would result in a leachate with a
pH of about 3.5 and elevated levels of some heavy metals.

In managing the reject material area Clarence has worked to:


• minimise the permeability of the material by the mixing of fine and coarse coal
reject materials;
• reduce permeability further through mechanical compaction;
• achieve the desired level of mechanical compaction through the use of a cone
penetrometer which measures compaction;
• prepare the annual extension to the reject emplacement area in December for the
next 12 months, with transfer of native tree seed stock onto progressively
rehabilitated areas, thus maximising the change of natural rehabilitation with the
summer rainfall. (Note: the preparation of the reject emplacement area involves
the immediate placement of cleared vegetation and soil onto an area to be
rehabilitated); and
• treating residual leachate, from the reject coal area generated during the
operational phase of mining, through the mine water treatment plant.
Benefits
• reduced quantities of leachate requiring treatment;
• maximising compaction to reduce the amount of disturbed land;
• reducing the long term requirement for management; and
• reducing the risk of spontaneous combustion.

64
To further reduce the residual leachate coming from reject emplacement areas and
potential groundwater contamination, Clarence is examining the feasibility of mixing
the coal washery reject wastes with a benign fine clay residue from the local
sandmine. This will provide multiple benefits in;

• reducing acid generation in the rejects;


• reducing the quantity of clay fines to be managed as tailings; and
• allowing the rejects/clay mixture to be deposited in the sand mine voids.

This latter benefit will provide a stable foundation for rehabilitation of the sand mine
while reducing the disturbed area necessary for placement of future coal washery
rejects.

CASE STUDY 8
PT KALTIM PRIMA COAL, KALIMATAN, INDONESIA

Introduction

PT Kaltim Prima Coal (KPC), an Indonesian incorporated company owned by British


Petroleum Co Plc of the UK and RTZ-CRA of Australia and the UK, has mined coal
at Sangatta since 1992. The mine is approximately 200 km north of Balikpapan on the
east coast of Kalimantan in Indonesia. The company currently mines 11.5 million
tonnes annually using trucks and shovels in a conventional open cut operation. The
climate at KPC is equatorial with high seasonal rainfall and low wind velocities.
Rainfall averages 2 500 millimetres annually and falls mainly as tropical
thunderstorms.

Acid mine drainage was detected at the mine in its second year of operation. Since
then, significant resources have been invested by the company to understand the
nature of the problem and to integrate acid control in the mining process.

Photo: PT Kaltim Prima Coal


Pt Kaltim Prima Coal, Kalimantan, Indonesia. Sampling spoils from blast holes.

65
Strategies for acid control

KPC's strategy for controlling acid drainage focuses on selective emplacement of


acid-forming material and the exclusion of oxygen from overburden sulphide, thereby
stopping pyrite oxidation. Several conditions exist where potential acid generation has
to be controlled. These include:

• existing dumps constructed prior to the understanding of the acid generation


problem;
• new dumps being built as a part of the current operations; and
• exposed coal floors rich in pyritic material during the mining operation.

Each requires a different approach.

Controlling acid production from surfaces of existing dumps

Prior to identifying the acid production potential at the mine, some early dump
surfaces were covered with acid producing overburden. All early dump surfaces have
been sampled to ascertain the chemical nature of their external layers and, where acid-
producing material has been identified, the surface has been encased with non reactive
material.

Research on cover designs has shown that the oxygen barrier has to achieve
permeability of less than 10-8 ms-1 with less than 10% air voids. Under the conditions
prevalent at KPC, this is achieved by either:

• placing 2.0 metres of non acid-forming overburden and 0.5 metres of topsoil
over pyritic material; or
• compacting half a metre of clay to the engineering criteria stated above,
followed by covering with 0.5-1.0 metre of topsoil to foster rehabilitation.

Strict engineering and geotechnical procedures must be followed to achieve adequate


clay compaction for oxygen exclusion.

New dumps and selective placement of overburden

Present day dumps are constructed with the aim of preventing acid drainage
occurring. This is achieved by ensuring that all highly reactive overburden is buried in
the core of dumps, away from a fluctuating water table and oxygen.

With knowledge gained from the overburden geochemistry study and local
geotechnical constraints, KPC has produced guidelines for constructing waste dumps
at its minesite.

The objectives of the guidelines are to:

• achieve long-term geotechnical stability of dumps;


• build dumps that resist erosion scour during construction and rehabilitation; and
• exclude potentially acid-forming rocks from dump surfaces.

66
To achieve these objectives it is imperative to emplace acid-forming material
selectively under non acid-forming material. For medium to long term stability all
materials in the outer 10 metres of the dumps must be non acid-forming. A typical
cross section of dumps at KPC is shown in Fig 17.0.

An important practice is that final drainage design works stabilise the surfaces and
minimise any future exposure of acid producing materials.

Photo: Kaltim Prima Coal


Loading spoils for selective placement.

Photo: Kaltim Prima Coal


Selective spoils placement at Kaltim Prima Coal.

Interim dumps

An intermediate dump face is an exposed face that will be over-dumped at a later


stage. It includes any dump face that is exposed for two months or less.

Interim overburden management specifications are as follows:

• any material that is highly acid-forming should be buried deep within the dump;
• such material should be placed in prepared cells within the dump, below the
projected long-term water table where it will not be subjected to fluctuating
water levels;
• non acid-forming overburden should form the outer five metres of intermediate
dump faces; and
• top surfaces of intermediate benches should be covered with at least one metre
of non acid-forming material.

67
Coal floors

Pyritic materials at KPC are mainly associated with coal roofs and floors. While roof
materials are removed as part of the mining process and placed within the core of the
dumps, floor material could be left exposed for lengthy periods. To prevent the
production of acid, floors are either dug out and placed in cores of dumps along with
coal roofs, or covered according to the engineering specifications listed above.

In a limited number of cases, acid is produced from exposed coal floors prior to
remedial action being implemented. Currently, such acid water from coal floors is
collected and treated with lime to achieve acceptable water chemistry prior to final
discharge. Two portable acid neutralisation plants are intermittently used for this
purpose. The plants have been very useful because of their cost effectiveness and their
mobility, as they can be moved to particular locations to provide short to medium
term treatment of acid water. A passive wetland biological filter to polish discharge
water is planned to be constructed in 1997.

Photo: PT Kaltim Prima Coal


Portable water treatment plant.

Figure 17.0 A typical dump profile showing selective


placement of mine spoil

Rehabilitation

Fundamental to any post-mining land use is re-establishment of vegetation. The


effective control of acid formation is critical to any rehabilitation program. Some
post-mining land uses being considered by KPC are production forestry, agriculture
and recreation.

68
More than 450 000 trees have been planted on 450 hectares on KPC's leases since
mining commenced. In 1996, the target was 206 hectares, an area exceeding that
cleared for mining for the first time. This trend will continue and areas of 200-400
hectares are planned to be rehabilitated each year. A detailed description of the
rehabilitation program is provided in Michaelsen, D and Macmillan, S (1996) under
Further Reading.

Conclusions

Acid drainage is endemic in East Kalimantan, associated not only with mining, but
also with any earthworks exposing pyritic materials. It is important to manage acid
drainage because of its deleterious effects on water chemistry, animals and vegetation.
In the mining context, perhaps the most significant impact is that acid-producing
surfaces will not revegetate.

A number of lessons have been learnt from KPC's experiences with acid drainage:

• acid prevention has to be integrated into the mining process; (Ideally, the control
program begins at the mine exploration stage when valuable baseline
information about potential acid formation can be obtained at relatively low
cost. Waste rock analysis is essential.)
• the stratigraphically controlled occurrence of acid-forming materials in steeply
dipping rocks, coupled with a mining method based on horizontal benches, has
complicated the accurate modelling of net acid generation at the mine; (The
approach taken by KPC has, however, provided adequate information to
manually schedule materials for selective placement within dumps. The
strategies described in this case study have shown to be cost-effective.)
• acid production must be prevented rather than treated; (If the acid-producing
potential of overburden is understood before the mining commences, the
problem can be arrested before it becomes costly. Selective placement of spoils
has proven an effective way to control acid production from open cut mining
operations.) and
• acid control strategies must be implemented completely and in a timely manner
to be effective. (A partially implemented acid control program is non
effectual—water quality continues to deteriorate and rehabilitation will not
succeed.)

In summary, experience at KPC suggests that it will be significantly more cost


effective to prevent the problem than trying to remedy it. KPC is continually striving
to improve its environmental performance by applying best practice approaches to the
prevention of acid drainage.

69

You might also like